Neuropharmacology 47 (2004) 1117–1134 www.elsevier.com/locate/neuropharm
Review
Phenotypic analysis of dopamine receptor knockout mice; recent insights into the functional specificity of dopamine receptor subtypes Andrew Holmes a,, Jean E. Lachowicz b, David R. Sibley c a
c
Section on Behavioral Science and Genetics, National Institute of Alcoholism and Alcohol Abuse, National Institutes of Health, Building 10, Room 3C217, Bethesda, MD 20892, USA b CNS/CV Biological Research, Schering Plough Research Institute, Kenilworth, NJ 07033, USA Molecular Neuropharmacology Section, National Institute of Neurological Disorders and Stroke, National Institutes of Health, Bethesda, MD 20892, USA Received 7 May 2004; received in revised form 20 June 2004; accepted 28 July 2004
Abstract The functional specificity of dopamine receptor subtypes remains incompletely understood, in part due to the absence of highly selective agonists and antagonists. Phenotypic analysis of dopamine receptor knockout mice has been instrumental in identifying the role of dopamine receptor subtypes in mediating dopamine’s effects on motor function, cognition, reward, and emotional behaviors. In this article, we provide an update of recent studies in dopamine receptor knockout mice and discuss the limitations and future promise of this approach. Published by Elsevier Ltd. Keywords: Dopamine; Receptor; Gene; Behavior; Knockout; Mouse
1. Introduction 1.1. Dopamine receptor subtypes and dopamine receptor ‘‘knockout’’ mice The brain dopamine system is organized into four anatomically distinct pathways (Lindvall and Bjorklund, 1978). Dopamine neurons in the substantia nigra project to the striatum to form the ‘‘nigrostriatal’’ pathway. A second major group of dopamine-containing neurons project from the ventral tegmental area to form the ‘‘mesolimbic’’ (innervating the nucleus accumbens, septum, olfactory tubercle, amygdala, and piriform cortex) and ‘‘mescocortical’’ (innervating medial prefrontal, cingulate and entorhinal cortices) pathways. A fourth, ‘‘tuberoinfundibular’’, pathway sends efferents from the arcuate nucleus of the hypothalamus to the intermediate lobe of the pituitary and the hypophyseal portal vessels of the median eminence. While these pathways are each associated with particular neural Corresponding author. Tel.: +1-3014-203519; fax: +1-3014801952. E-mail address:
[email protected] (A. Holmes).
0028-3908/$ - see front matter Published by Elsevier Ltd. doi:10.1016/j.neuropharm.2004.07.034
functions and disease states, such functional demarcation is far from absolute. Further separation and refinement of dopaminergic function is achieved, in part, via actions at different receptor subtypes. Dopamine mediates its neural effects via actions at both presynaptic and postsynaptic dopamine receptors. Five separate dopamine receptor subtypes have been identified, all belonging to the seven transmembrane G-protein coupled receptor family (Civelli et al., 1993; Gingrich and Caron, 1993; Sibley et al., 1993). Based on its pharmacological and signaling properties, the D1R-like subfamily, comprising D1R and D5R subtypes, is differentiated from the D2R-like subfamily, comprising D2R, D3R and D4R subtypes (Kebabian and Calne, 1979; Missale et al., 1998). D1R-like receptors stimulate signal transduction by coupling to Gs proteins and subsequent activation of adenylyl cyclase and cAMP production. D2R-like receptors couple to Gi/o-like proteins and suppress signal transduction via inhibition of adenylyl cyclase and cAMP production and modulation of ion channels. The divergent intracellular effects of dopamine receptors, together with their divergent neuroanatomical
1118
A. Holmes et al. / Neuropharmacology 47 (2004) 1117–1134
localization, strongly suggest that individual dopamine receptor subtypes mediate distinct functional properties of dopamine. There is no doubt that elucidating this functional specificity would represent a major advance in our understanding of how dopamine governs neural function and impacts a variety of debilitating neurological and neuropsychiatric diseases. Unfortunately, the majority of available dopamine receptor agonists or antagonists do not act with specificity at individual receptor subtypes within the D1R-like or D2R-like subfamilies, thereby limiting their utility as research tools. There are currently no ligands with greater than 10-fold selectivity for D1R vs. D5R. Within the D2R-like family, there are compounds with >100-fold selectivity for D4R vs. D2R, or D3R vs. D2R, but still no D2R-selective agonist and antagonists. D2RL and D2RS isoforms of the D2R cannot be selectively targeted with pharmacological agents. Gene targeting techniques in the mouse have evolved as a widely used approach to elucidate the functions of specific molecules found in the brain (Crawley, 2000; Bucan and Abel, 2002). Given the aforementioned limitations of a traditional behavioral pharmacological approach to study dopamine receptor subtype function, the ability to functionally ‘‘knockout’’ a dopamine receptor with great specificity has made it an attractive and fruitful strategy. Several authors have reviewed the many studies that have reported on neural and behavioral phenotypes in dopamine receptor knockout (KO) mice (Sibley, 1999; Glickstein and Schmauss, 2001; Waddington et al., 2001; Zhang and Xu, 2001; Tan et al., 2003; Viggiano et al., 2003). The goal of the present article is to provide an update of recent research in this rapidly moving field and attempt to place these developments within the context of prior findings.
2. D1R KO mice; movement, motivation and memory 2.1. General D1R KO mice were the first dopamine receptor KO mice to be generated. Two lines of D1R KO mice exist (Drago et al., 1994; Xu et al., 1994b). D1R KO mice show normal appearance and no obvious neurological defects, but exhibit growth retardation and low survival after weaning (Drago et al., 1994; Xu et al., 1994a, b). This failure to thrive can be rescued by providing KO mice with easy access to a palatable food, such as wet chow on the cage floor, suggesting that it may relate to a deficit in fine motor control (Drago et al., 1994; Xu et al., 1994a, b; see also Section 2). This contrasts with the apparent loss of the motivation to feed that is observed in mutant mice
with virtually no dopaminergic neurotransmission (Cannon and Palmiter, 2003). The D1R is abundant in the striatum, amygdala and olfactory tubercle, and is also found in the hippocampus, globus pallidus, hypothalamus, septum, substantia nigra, and throughout the cerebral cortex (Levey et al., 1993; Ariano and Sibley, 1994). Radioligand binding of D1R-like receptors is negligible in the brain of D1R KO mice, although some binding, presumably to D5R, is detectable in hippocampus (Friedman et al., 1997; Becker et al., 2001; Montague et al., 2001; Wong et al., 2003a). D2R-like binding and dopamine transporter binding also appear normal in D1R KO mice (Xu et al., 1994a; El-Ghundi et al., 1999; but see Wong et al., 2003a). However, there are fewer total dopamine (tyrosine hydroxylase-positive) neurons in D1R KO mice than wild type (WT) controls (Parish et al., 2002). In addition, midbrain levels of dopamine are elevated and dopamine metabolites are altered in D1R KO mice, suggesting that there may be a compensatory increase in dopamine synthesis and/or release following inactivation of the receptor (El-Ghundi et al., 1998; Parish et al., 2002). Lastly, striatal levels of dynorphin and substance P, both of which are synthesized in D1Rexpressing neurons, are reduced in D1R KO mice (Drago et al., 1994; Xu et al., 1994b; Wong et al., 2003a). Psychostimulant-induced increases in dynorphin, but not substance P, are also abolished in these mice (Drago et al., 1996; Moratalla et al., 1996). 2.2. Motor functions While D1R-like antagonists profoundly reduce locomotor activity, several laboratories have found that D1R KO mice exhibit increased locomotor activity and retarded locomotor habituation in an open field apparatus (Xu et al., 1994a, b, 2000b; Crawford et al., 1997; Clifford et al., 1998; Karasinska et al., 2000; Becker et al., 2001; Miyamoto et al., 2001; Centonze et al., 2003a; McNamara et al., 2003; Wong et al., 2003b). Additionally, D1R KO mice have been reported to show reduced orofacial movements and impaired motor coordination on the rotarod test, as well as reduced spontaneous grooming behavior and an absence of novelty- and neuropeptide-induced grooming (Cromwell et al., 1998; Drago et al., 1999; Karasinska et al., 2000; Tomiyama et al., 2002). The hyperactivity- and hypoactivity-inducing effects of D1R-like (but not D2R-like) agonists and antagonists are abolished in D1R KO mice (Xu et al., 1994a, b; Drago et al., 1996; Moratalla et al., 1996; Clifford et al., 1999; Miyamoto et al., 2001; Tomiyama et al., 2002; Centonze et al., 2003a; McNamara et al., 2003). In the context of dopamine’s putative role in schizophrenia (see Section 4.5), it is noteworthy that the hyperlocomotor-inducing effects of the psychotomimetic NMDA
A. Holmes et al. / Neuropharmacology 47 (2004) 1117–1134
antagonist, ketamine, are also lost in the D1R KO mice (Miyamoto et al., 2001). Taken together, these findings suggest an important role for D1R in maintaining spontaneous motor behaviors, and appear to affirm the relative importance of D1R over D5R in mediating the motor effects of D1R-like agonists (see also Section 3.2). However, baseline hyperactivity, motor incoordination and absent locomotor responses to D1R-like agonists have not been found in all studies of D1R KO mice, and in some cases opposite (e.g., a hypoactive baseline), phenotypes have been observed (Drago et al., 1996; Smith et al., 1998; Clifford et al., 1999; El-Ghundi et al., 1999; Karasinska et al., 2000). The reason for these inconsistencies is not yet fully understood. One widely discussed possibility is that analysis of motoric phenotypes in D1R KO mice may have been confounded by the genetic background on which the mutation was present, such that genes contributed by one of the parental strains (i.e., a 129 substrain), rather than the loss of D1R, may underlie the observed locomotor abnormalities (for a fuller discussion, see Section 4.3). However, this explanation appears to be discounted by the recent finding that locomotor phenotypes, such as reduced locomotor habituation, were actually more prominent in D1R KO mice backcrossed onto a congenic (C57BL/6) genetic background than in mice on a mixed genetic background (McNamara et al., 2003). This study does nonetheless strongly emphasize the critical influence of genetic background on behavioral phenotypes in dopamine receptor KO mice. Another factor influencing variability in D1R locomotor phenotypes across studies is likely to stem from inter-laboratory differences in the methods used to assess these behaviors. It is increasingly accepted that variations in task parameters such as the size, material, shape, and illumination of the apparatus used, the test duration employed, and the prior experimental experience of the subject can strongly influence a variety of mouse behaviors, particularly in tasks based on exploratory locomotion (Holmes and Rodgers, 1998; Crabbe et al., 1999; Holmes, 2001; McIlwain et al., 2001). In this context, there is growing evidence that the dopamine system is acutely sensitive to psychological stressors (Arnsten and Goldman-Rakic, 1998). Thus, ostensibly subtle variations in task protocol and procedure, as well rearing and husbandry conditions, may produce marked variation in the level of stress/ dopamine challenge evoked by a behavioral task and, in doing so, modify the manifestation of abnormal D1R KO phenotypes (Clifford et al., 2000). This scenario may apply equally to the study of behavior in other dopamine receptor KO mice. Unfortunately, there are no easy solutions to this problem. The adoption of ‘‘standardized’’ methods of behavioral analysis is likely to be impractical beyond a
1119
rudimentary level (for relevant discussion, see Wahlsten et al., 2003). A valuable step forward would be to better understand the relationship between stress and the dopamine system. In addition, some authors have proposed that refining behavioral analysis may clarify the precise nature of phenotypic alterations in dopamine receptor KO mice (for review see Waddington et al., 2001). Indeed, the adoption of more sophisticated methods of assessment, that incorporate ethological measures of distinct components of the behavioral repertoire (Rodgers et al., 1997), has proved valuable to the phenotypic description of locomotor phenotypes in D1R (and other dopamine receptor) KO mice (Drago et al., 1996; Clifford et al., 1999). 2.3. Reward D1R KO mice have been instrumental in clarifying the role of the D1R subtype in dopaminergic mediation of reward-related behaviors. Consistent with the effects of D1R-like antagonist treatment (Kalivas and Stewart, 1991), deletion of the D1R KO either eliminates or markedly attenuates the hyperactivity-inducing and behavioral-sensitizing effects of acute and chronic treatment with psychostimulants or morphine (Drago et al., 1996; Xu et al., 1996, 2000b; Crawford et al., 1997; Becker et al., 2001, but see Karper et al., 2002). Psychostimulant treatment in D1R KO mice also fails to inhibit firing or induce expression of brain derived nerve growth factor (BDNF) and other genes in brain regions implicated in reward, such as the nucleus accumbens (Xu et al., 1994a, 2000b; Miner et al., 1995; Moratalla et al., 1996; Crawford et al., 1997; Zhang et al., 2002a,b), an issue which we return to later. Recent studies have complimented these observations by demonstrating that psychostimulant-induced phosphorylation of DARPP-32, an intracellular signaling molecule thought to mediate the neural effects of psychostimulants, is also compromised in D1R KO mice (Svenningsson et al., 2003). However, reward-related behaviors in D1R KO mice have yet to be thoroughly studied. An early investigation found that D1R KO mice demonstrated normal responses to the rewarding effects of cocaine in the conditioned place preference test (Miner et al., 1995). More recently, D1R KO mice have been found to show reduced voluntary ethanol consumption, as well as reduced operant responding for a sucrose reinforcer (El-Ghundi et al., 1998, 2003). The possibility, suggested by these findings, that deletion of the D1R may attenuate the reinforcing properties of rewarding stimuli, would be consistent with emerging clinical data; e.g., the euphoric effects of cocaine can be blunted by a D1R-like antagonist in cocaine addicts (Romach et al., 1999). Given the clinical implications of identifying novel dopamine receptor-based therapeutic targets for
1120
A. Holmes et al. / Neuropharmacology 47 (2004) 1117–1134
addiction, further studies of reward-related behavior in D1R KO mice will be of considerable importance. 2.4. Cognition Dopamine is known to play an important role in cognition (Robbins, 2000). Interestingly, a number of studies have reported deficits on hippocampal-mediated behavioral tasks, such as the Morris water maze, in D1R KO mice (El-Ghundi et al., 1998; Smith et al., 1998; Karasinska et al., 2000). Moreover, D1R KO mice exhibit impaired long-term potentiation, a molecular correlate of learning and memory, in both hippocampal and corticostriatal neurons (Matthies et al., 1997; Centonze et al., 2003a). However, performance on other hippocampal-mediated tests such as contextual fear conditioning and passive avoidance appears normal in D1R KO mice, and more work is required to ascertain the precise nature of hippocampal function in these mice (El-Ghundi et al., 1999, 2001). A growing literature implicates the prefrontal cortex in the mediation of behaviors including working memory and cognitive flexibility, response inhibition, and extinction of learned fear responses (Goldman-Rakic, 1995; Milad and Quirk, 2002; Chudasama and Robbins, 2003). In this context, recent findings demonstrating deficits in fear extinction and reversal learning in D1R KO mice suggest that prefrontal cortex function may be impaired (El-Ghundi et al., 1999, 2001, 2003). This is supported by the demonstration that long-term potentiation in prefrontal cortex neurons is abnormal in D1R deficient mice (Huang et al., 2004). Indeed, a prefrontal deficit in D1R KO mice would also be highly congruent with (1), the rich localization of D1R in the prefrontal cortex (Levey et al., 1993; Ariano and Sibley, 1994), (2), D1R-like modulation of prefrontal cortex synaptic plasticity (Otani et al., 1998; Gurden et al., 2000; Huang et al., 2004) and (3), evidence that D1R-like agonists and antagonists administered directly into the prefrontal cortex affect cognitive performance in rodents and non-human primates (Williams and Goldman-Rakic, 1995; Granon et al., 2000; Robbins, 2000; Lidow et al., 2003). Thus, the existing data suggest that D1R KO mice could provide an interesting model to study the role of this subtype in prefrontal cortex-mediated behaviors.
3. D5R KO mice; still an enigma? 3.1. General D5R KO mice are viable, healthy and develop without the growth retardation seen in D1R KO mice (Holmes et al., 2001; Hollon et al., 2002). Moreover, despite evidence that antisense knockdown of the D5R
in the ventromedial hypothalamus inhibits lordosis in female rats (Apostolakis et al., 1996), D5R KO mice are fertile and reproduce normally. D5R KO mice do, however, develop abnormally high blood pressure (Hollon et al., 2002) and show increased vulnerability to cysteamine-induced gastric lesions (Hunyady et al., 2001). Across species, the neural expression of the D5R is distinct from and more limited than D1R; in the rodent brain, the D5R is found in the hippocampus, prefrontal cortex and basal ganglia, as well as thalamic and hypothalamic nuclei (Ciliax et al., 2000). Confirming the preponderance of the D1R relative to the D5R in the mouse brain, D1R-like autoradiographic binding in D5R KO mice is almost indistinguishable from WT controls (Montague et al., 2001; Hollon et al., 2002; Laplante et al., 2004). To date, possible changes in brain D2R-like receptors in D5R KO have not been demonstrated. 3.2. Motor functions Consistent with a relatively specialized role for the D5R, phenotypic analysis of D5R KO mice has revealed fewer behavioral abnormalities than observed in D1R KO. Basal locomotor activity, motor coordination, prepulse inhibition, and anxiety-like behaviors all appear normal in D5R KO mice (Holmes et al., 2001; Elliot et al., 2003). Operant responding for cocaine was also recently shown to be normal in these mice (Elliot et al., 2003). Preliminary data have indicated a gender-specific antidepressant-like phenotype in D5R KO mice tested in a mouse model of ‘‘behavioral despair’’ (Holmes et al., 2001). If robust, this phenotype may provide a novel insight into the role of D5R in stress-related behavior, and could relate to the aforementioned localization of D5R in the hippocampus, a brain region implicated in depression (MacQueen et al., 2003). The hyperactivity-inducing effects of D1R-like agonist and, in a separate study, cocaine, were found to be mildly attenuated in D5R KO (Holmes et al., 2001; Elliot et al., 2003). These findings extend studies showing a near complete loss of the same effects of D1R-like agonists in D1R KO mice (see Section 2). Because there is no evidence of altered D1R binding in D5R KO mice, the most parsimonious explanation for these findings is that D5R normally contributes to the hyperactivity-inducing effects of dopamine agonists (Centonze et al., 2003a). This hypothesis is supported by other recent findings. First, while dopaminergic excitation of striatal interneurons is blocked by treatment with a D1R-like antagonist in WT mice, this response is retained in D1R KO mice (Centonze et al., 2003b). Second, D1R-like agonists are still able to modulate neuronal activity in the subthalamic nucleus
A. Holmes et al. / Neuropharmacology 47 (2004) 1117–1134
(a major region controlling basal ganglia output) in the absence of the D1R (Baufreton et al., 2003). These emerging findings have led to the proposal that D1R and D5R exert synergistic control over locomotor activity when the subtypes are expressed on the same striatal neuronal type (medium spiny) via a common signal transduction mechanism (Centonze et al., 2003a). Ultrastructural differentiation of D1R and D5R within corticostriatal neurons also raises the possibility of mutually opposing actions of the subtypes (Bergson et al., 1995; Centonze et al., 2003b). 3.3. Cognition A notable exception to the higher expression of the D1R vs. D5R in the brain is in the hippocampus, where the pattern is reversed (Ciliax et al., 2000). Recent research in D5R KO mice has begun to shed light on the function of hippocampal D5R. D5R KO mice display reduced extracellular levels of acetylcholine in the hippocampus as well as significantly blunted acetylcholine-releasing responses to D1R-like agonist treatment in this region (Laplante et al., 2004). These findings concur with previous evidence indicating that the antisense knockdown of hippocampal D5R attenuates D1R-like agonist stimulation of hippocampal acetylcholine release in rats (Hersi et al., 2000). Given the well-established role of cholinergic neurotransmission in mediating learning and memory (Everitt and Robbins, 1997), combined with evidence of dopamine D1R-like facilitation of hippocampal long-term potentiation and memory function (Bach et al., 1999; Li et al., 2003), these findings suggest an important modulatory role for the D5R in hippocampal-mediated cognition. However, to date, behavioral studies have found little evidence of impaired learning and memory on various hippocampal-mediated tasks in D5R KO mice (Holmes et al., 2001). More detailed studies of cognitive function in D5R KO mice are awaited.
4. D2R; splice variants enrich the picture 4.1. General Three separate lines of D2R KO mice have been generated (Baik et al., 1995; Kelly et al., 1997; Jung et al., 1999). To date, these mice have been the most well studied of the dopamine receptor KO mice. Consistent with the D2R’s role as an inhibitory mediator of pituitary hormone synthesis and secretion (Sibley and Creese, 1983) D2R KO mice exhibit reduced pituitary growth hormone release, develop pituitary tumors and, as a result of elevated glucocorticoid levels, adrenal hypertrophy (Kelly et al., 1997; Saiardi et al., 1997; Saiardi and Borrelli, 1998; Asa et al., 1999; Diaz-Torga
1121
et al., 2002). As pituitary hormone function is normal in D2RL KO mice (see Section 4.6), D2R mediation of this function appears to occur through the D2RS isoform (Xu et al., 2002). D2R KO mice also show increased blood pressure and susceptibility to sodiuminduced hypertension (Li et al., 2001, 2002; Ozono et al., 2003; Ueda et al., 2003). Otherwise, and despite some reports of growth retardation, D2R KO mice appear healthy. The D2R is expressed at much higher levels in brain than the other D2R-like subtypes. The D2R is especially prominent in the vicinity of the mesencephalic dopamine cell body regions, as well as in the dorsal and ventral striatum and the olfactory tubercle (Meador-Woodruff et al., 1989). The loss of the D2R in KO mice is associated with reduced levels of neurotrophins and fewer total striatal dopamine (i.e., tyrosine hydroxylase-positive) neurons (Bozzi and Borrelli, 1999; Parish et al., 2002). In addition, D2R KO mice exhibit significantly increased striatopallidal enkephalin and reduced striatonigral substance P mRNA, as well increased cytochrome oxidase expression in the subthalamic nucleus (Murer et al., 2000). D2R KO mice show a transient postnatal increase in D3R, and in some reports, a decrease in D1R-like binding (Baik et al., 1995; Xu et al., 1997; Kelly et al., 1998; Jung et al., 1999). Although the motoric effects of D1R-like agonists and antagonists are largely normal in D2R KO mice (Kelly et al., 1998; Phillips et al., 1998; Boulay et al., 1999a; Tomiyama et al., 2004), D1R-like agonist-induced stereotypical behavior is exaggerated (Glickstein and Schmauss, 2001, 2004). Moreover, D2R KO mice show significantly blunted striatum and prefrontal cortex activation (as assayed by c-fos expression) in response to D1R-like agonists (Jung et al., 1999; Schmauss, 2000; Schmauss et al., 2002; Glickstein and Schmauss, 2004). These alterations in D2R KO cortical responsivity are associated with deficits in working memory performance, and can be reversed by treatment with methamphetamine (Schmauss, 2000; Glickstein et al., 2002). Parenthetically, transgenic overexpression of D1R in extrastriatal regions does not significantly alter behavioral abnormalities in D2R KO mice (Dracheva et al., 1999; Dracheva and Haroutunian, 2001). Taken together, the available evidence indicates certain D1R functional alterations in D2R KO mice, although the precise nature of these alterations is yet to be fully elucidated. 4.2. Autoreceptor function D2R-like receptors are localized both postsynaptically and presynaptically on dopamine terminals and soma. Consistent with a major functional role of the D2R in regulating dopaminergic neuronal activity (Roth, 1984), the ability of dopamine to inhibit the fir-
1122
A. Holmes et al. / Neuropharmacology 47 (2004) 1117–1134
ing of neurons in the substantia nigra or ventral tegmental area, or inhibit dopamine release in striatal projections areas, is lost in D2R KO mice (Mercuri et al., 1997; Rouge-Pont et al., 2002). In addition, deletion of the D2R (but not D3R) abolishes the ability of D2Rlike agonists to inhibit striatal dopamine synthesis and release, activate G-protein-gated inwardly rectifying potassium channels and increase locomotor activity (Kelly et al., 1998; Koeltzow et al., 1998; L’Hirondel et al., 1998; Boulay et al., 1999a, 2000; Benoit-Marand et al., 2001; Clifford et al., 2001; McNamara et al., 2002; Schmitz et al., 2002; Davila et al., 2003; Lindgren et al., 2003; Tomiyama et al., 2004). Thirdly, the capacity of striatal dopamine neurons to attenuate amphetamine-induced dopamine release is disrupted in D2R KO mice (Schmitz et al., 2001). Taken together, these findings confirm that the D2R serves as the principal autoreceptor on dopamine neurons. However, despite the loss of this function, neither synaptic dopamine levels nor tissue concentrations of dopamine or its metabolites are markedly altered in D2R KO mice (Benoit-Marand et al., 2001; Rouge-Pont et al., 2002; but see Koeltzow et al., 1998; Jung et al., 1999; Parish et al., 2002). This profile indicates that either the D2R does not exert strong tonic control of dopamine activity or, more likely, that compensatory mechanisms have fulfilled this role in D2R KO mice. In this context, there is some evidence that the other major regulator of extracellular striatal dopamine, the dopamine transporter, is upregulated in the striatum of D2R KO mice (Parish et al., 2002). Functional changes in dopamine transporter-mediated striatal dopamine reuptake have also been found in D2R KO mice, although the magnitude and direction of the changes have varied across studies (Dickinson et al., 1999; Schmauss et al., 2002). 4.3. Motor functions D2R KO mice exhibit significantly lower levels of locomotor activity than WT controls (Baik et al., 1995; Kelly et al., 1998; Boulay et al., 1999a, 2000; Jung et al., 1999; Aoyama et al., 2000; Clifford et al., 2000; Zahniser et al., 2000; Dracheva and Haroutunian, 2001; Chausmer et al., 2002; McNamara et al., 2002; Tran et al., 2002; Vallone et al., 2002; Palmer et al., 2003). Moreover, profound deficits in gait, posture and motor coordination in D2R KO mice suggested a fundamental role for the receptor in mediating dopaminergic control of these behaviors (Baik et al., 1995; Jung et al., 1999; Aoyama et al., 2000). However, gait and motor coordination deficits in D2R KO mice have not been found in all investigations. As noted in Section 2.2, genetic background has been found to be a significant influence on motor phenotypes in D1R KO mice. There are marked strain
differences in motor performance on the rotarod, open field and other motor tasks (Tarantino et al., 2000; Holmes et al., 2002). Therefore, the fact that the motor incoordination phenotype in D2R KO mice resembles that of the 129 parental strain is consistent with the possibility that this phenotype is a false-positive caused by 129 parental genes (Crusio, 1996; Gerlai, 1996). This explanation is supported by the finding that repeated backcrossing of the D2R null mutation onto a C57BL/6 genetic background effectively ‘‘rescued’’ motor coordination deficits in these mice (Kelly et al., 1998). An alternative possibility is that when the mutation is backcrossed into the C57BL/6J genetic strain, epistasis interactions between the null mutation and C57BL/6J background gene(s) protect against motor coordination deficits in D2R KOs (for discussion, see Holmes et al., 2003). Irrespective of the underlying reasons, the significant influence of genetic background on both D2R KO and D1R KO locomotor phenotypes strongly advises the use of congenic backgrounds in future studies of these and other dopamine receptor KO mice. There is growing evidence of the importance of functional interactions between the D2R and A2A adenosine receptor subtype in the pathophysiology and treatment of dopamine-related movement disorders, such as Parkinson’s disease (Kase et al., 2003). The D2R KO mouse has recently proven to be a useful tool for studying D2R–A2AR interactions. While A2AR binding density is normal in D2R KO mice, there is evidence of functional uncoupling of the A2AR in these mice (Aoyama et al., 2000; Zahniser et al., 2000). Behaviorally, the hyperactivity-inducing effects of A2AR antagonists are reduced, but not completely abolished, in D2R KOs (Aoyama et al., 2000; Zahniser et al., 2000; Chen et al., 2001). These and related findings have led to the hypothesis that the D2R and A2AR likely exert their effects on motor function via opposing, yet semi-independent mechanisms, and encourage further delineation of these interactions in search of novel therapeutic approaches to the treatment of movement disorders. 4.4. Reward Reward-related responses to a number of drugs of abuse are abnormal in D2R KO mice. Independent studies have shown that voluntary ethanol consumption, ethanol-induced conditioned place preference, operant responding for ethanol, ethanol-induced sedation and ataxia are all reduced in D2R KO mice, while the locomotor stimulating and sensitizing effects of ethanol are increased (Phillips et al., 1998; Cunningham et al., 2000; Risinger et al., 2000; Palmer et al., 2003). Extending these findings and demonstrating that the influence of genetic background on D2R KO
A. Holmes et al. / Neuropharmacology 47 (2004) 1117–1134
behavioral phenotypes extends to ethanol responses (see Section 4.3), a recent study has shown that augmented locomotor stimulant effects of ethanol are observed when the mutation is on a C57BL/6J, but not 129, genetic background, while reduced ethanolinduced ataxia occurs independently of background strain (Palmer et al., 2003). Notwithstanding, these data in D2R KO mice extend evidence that the mouse D2R (Drd2) is a major candidate gene underlying variation in ethanol-related behaviors (Phillips et al., 1994). D2R KO mice do not self-administer morphine and fail to develop conditioned place preference to morphine, while morphine-induced hyperactivity and morphine withdrawal (as well as brain l-opioid receptor levels) are normal in D2R KOs (Maldonado et al., 1997; Elmer et al., 2002; see also Smith et al., 2002). Further studies have suggested that the role of the D2R in opiate reward may be restricted to conditions in which the incentive to obtain the drug is especially high, as during dependence and withdrawal, while (Dockstader et al., 2001). Electrophysiological data also indicate that abnormal nucleus accumbens responses to reward in D2R KO mice may be highly specific to particular phases of reward processing and, more generally, that striatal synaptic plasticity is grossly abnormal in these mice (Calabresi et al., 1997; Tran et al., 2002). D2R KO responses to psychostimulants are similarly complex. The locomotor stimulant effects of cocaine are diminished in D2R KO mice (Chausmer and Katz, 2001; Chausmer et al., 2002), while selfadministration of moderate–high doses of cocaine is increased in D2R KO mice; an effect mimicked by D2R-, but not D3R- or D4R-, preferring antagonists (Caine et al., 2002). However, unlike normal mice treated with D2R-like antagonists, D2R KOs are fully able to discriminate cocaine from saline (Chausmer and Katz, 2001; Chausmer et al., 2002). These somewhat disparate findings concur with emerging clinical data pointing to a critical, but complex, role for D2R in reward and addition (Volkow et al., 2002). 4.5. Sensorimotor gating Startle responses to sudden high-intensity sensory stimuli (e.g., a loud noise) are inhibited when closely preceded by a low-intensity stimulus (a weaker sound). This ‘‘prepulse inhibition’’ (PPI) response is impaired in schizophrenia and certain other neuropsychiatric disorders (Braff et al., 2001), and has proven a valuable assay to delineate the actions of antipsychotics in mutant mouse models (Geyer et al., 2002). Rodent behavioral pharmacological data have established a significant role for D2R in mediating the PPI-impairing effects of psychotomimetic drugs such as amphetamine (Geyer et al., 2002). In agreement with these findings, D2R KO mice fail to show amphetamine-induced
1123
disruption of PPI (Ralph et al., 1999). Demonstrating the specificity of these effects, amphetamine-induced deficits were not found in D3R KO mice or D4R KO mice, and the PPI-deficits caused by treatment with a NMDA receptor antagonist were unaltered in D2R KO mice (Ralph et al., 1999; Ralph-Williams et al., 2002). Further work has demonstrated that amphetamine-induced disruption of PPI is not altered in D1R KO, while responses to D1R-like agonists are abolished in D1R KO and mildly attenuated in D5R KO mice (Holmes et al., 2001; Ralph-Williams et al., 2002). Thus, these data suggest that an essential role for the D2R in mediating sensorimotor gating deficits caused by amphetamine, but probably not all PPI-impairing drugs. In view of the long-standing hypothesis that the dopamine system is central to the pathophysiology and treatment schizophrenia, more work in this area utilizing dopamine receptor KO mice is expected. 4.6. Splice variants Across species the D2R mRNA is alternatively spliced to produce short (D2RS) and long (D2RL) isoforms, the latter containing an additional 29 amino acid insertion in the third cytoplasmic loop of the receptor (Giros et al., 1989; Mack et al., 1991). The 29 amino acid insertion sequence in the D2L isoform is encoded by an 87 bp cassette exon in the D2R gene (Monsma et al., 1989). In the mouse brain, the D2RL isoform is much more highly expressed and shows partially discrete neuroanatomical localization compared to the D2RS isoform, suggesting functional specificity (Mack et al., 1991; Neve et al., 1991). D2RL KO mice have been generated to explore this by selectively targeting the 87 bp exon which encodes the 29 amino acid sequence defining the D2RL isoform, thereby producing mice in which the D2R gene is only capable of generating the D2RS isoform (Usiello et al., 2000; Wang et al., 2000). D2R binding is normal in D2RL KO mice, presumably via increased D2RS binding (Wang et al., 2000). However, behavioral analysis has shown that D2RL KO mice exhibit a number of behavioral abnormalities. These phenotypes include diminished conditioned place preference to morphine, reduced aggression and impaired avoidance learning (Vukhac et al., 2001; Smith et al., 2002). D2RL KO mice also demonstrate reduced basal open field locomotor activity in some studies (Wang et al., 2000; but see Usiello et al., 2000; Xu et al., 2002; Fetsko et al., 2003). A more consistent finding has been that D2RL KO mice exhibit blunted hypoactive/cataleptic responses to D2R-like antagonists and reduced stereotypical responses to psychostimulants (Usiello et al., 2000; Wang et al., 2000; Xu et al., 2002; Fetsko et al., 2003). The baseline locomotor hypoactivity and the reduced
1124
A. Holmes et al. / Neuropharmacology 47 (2004) 1117–1134
responsivity to treatment with D2R-like ligands observed in D2RL KO mice resemble the phenotypic alterations found in D2R KO mice (see Section 4.3). However, there are clear phenotypic differences between D2R KO and also D2RL KO mice. D2R-like agonist-induced stereotypical behavior and D2R-like agonist activation of postsynaptic phosphoproteins, such as DARPP-32, which are both absent in D2R KO mice, are maintained in D2RL KO mice (Fetsko et al., 2003; Lindgren et al., 2003). Moreover, D2R autoreceptor-mediated inhibition of dopamine synthesis and release in response to D2R-like agonists or psychostimulants remain intact in D2RL KO mice (Usiello et al., 2000; Rouge-Pont et al., 2002). Lastly, in contrast to the loss of PPI-impairing effects of psychostimulants in D2R KO mice discussed in the preceding section, these responses are extant in D2RL KO mice (Xu et al., 2002). Taken together, these diverse data demonstrate that the D2RS and D2RL isoforms are responsible for discrete functions of the D2R. The available evidence suggests that D2RL is the isoform responsible for postsynaptic D2R effects, while D2RS may mediate D2R autoreceptor control of dopamine release. Clearly, such functional dissociation has major implications for developing novel pharmacotherapeutics targeting the D2R.
5. D3R; reward and cooperativity 5.1. General Three lines of mutant mice lacking functional D3R have been generated (Accili et al., 1996; Xu et al., 1997; Jung et al., 1999). D3R KO mice show normal appearance, growth, fertility, and no gross neurological dysfunctions, but develop renin-dependent hypertension (Asico et al., 1998; Jose et al., 1998). Brain binding density of other dopamine receptors, including the D1R and D2R, also appears normal in D3R KO mice (Accili et al., 1996; Xu et al., 1997; Wong et al., 2003a). While basal feeding and body weight are normal in D3R KOs, these mice deposit greater amounts of fat when maintained on a high-fat diet and show blunted hyperphagic responses to treatments such as leptin and insulin (Benoit et al., 2003; McQuade et al., 2004). These novel data suggest a potential specific role for D3R in body weight regulation. D2R-like receptors are known to be important regulators of body temperature, and D3R KO mice have helped clarify the role of D3R in these effects. Early pharmacological studies showing that hypothermia induction by mixed D2R/D3R agonists correlated highly with affinity for D3R implied that these effects were D3R-mediated (Millan et al., 1995). However, the hypothermia-inducing effects of D2R/D3R agonists are
normal in D3R KO mice but absent in D2R KOs, strongly suggesting that it is principally the D2R, rather than the D3R, that mediates these effects (Boulay et al., 1999a; Xu et al., 1999; Perachon et al., 2000). As discussed earlier, studies of D2R KO mice have provided evidence supporting the primacy of the D2R subtype as the major presynaptic autoreceptor on dopamine neurons (see Section 4.2). An auxiliary role for D3R in controlling dopamine release via postsynaptically mediated negative feedback has been proposed (Schwartz et al., 1993). While most measures of dopamine content, uptake and neuronal firing are normal in D3R KO mice, some reports note higher basal striatal extracellular dopamine (Xu et al., 1997; Koeltzow et al., 1998; but see Zapata et al., 2001). In addition, stereotypical and striatal responses (i.e., c-fos and dynorphin expression) to chronic cocaine treatment are exaggerated in D3R KOs (Xu et al., 1997; Carta et al., 2000). Lastly, a recent study has shown that D1R and D2R-mediated G-protein activation and calcium influx is enhanced in cultured forebrain neurons from D3R KO mice (Mizuo et al., 2004). Taken together, these findings are consistent with an inhibitory influence of D3R, at least in striatal neurons. However, there are also reports that locomotor responses to psychostimulants are unaltered in D3R KO mice and that their striatal responsivity to D1R-like agonist and chronic cocaine treatment is diminished rather than increased (Jung and Schmauss, 1999; Betancur et al., 2001). Moreover, while D3R-preferring agonists fail to inhibit striatal dopamine release in D3R KO mice, the locomotor effects of putatively selective D3R agonists and antagonists are maintained in D3R KO mice, but lost in D2R KOs, suggesting that the D2R rather than D3R is the principle receptor mediating these behavioral changes (Boulay et al., 1999a, b, 2000; Xu et al., 1999; Betancur et al., 2001; Zapata et al., 2001; Pritchard et al., 2003). 5.2. Locomotion, emotion and reward D3R KO mice exhibit locomotor hyperactivity in some studies, but not others (Xu et al., 1997; Boulay et al., 1999a,b, 2000; Depoortere, 1999; Jung and Schmauss, 1999; Jung et al., 1999; Carta et al., 2000; Karasinska et al., 2000; Betancur et al., 2001; McNamara et al., 2002; Vallone et al., 2002; Boyce-Rustay and Risinger, 2003; Pritchard et al., 2003; Wong et al., 2003b). A possible cause of this variability is that the increase in activity in D3R KOs may be transient and restricted to the early phase of open field testing (Accili et al., 1996; Xu et al., 1997; Vallone et al., 2002). One interpretation of an exaggerated locomotor response during initial exposure to a novel environment is that it indicates reduced neophobia/anxiety-like behavior
A. Holmes et al. / Neuropharmacology 47 (2004) 1117–1134
(Holmes, 2001). Supporting this possibility, some, but not all, studies D3R KO mice show a reduced anxietylike phenotype on the elevated plus-maze and open field (Steiner et al., 1997; Xu et al., 1997; Karasinska et al., 2000; Vallone et al., 2002). However, these data in KO mice are seemingly at odds with the anxiolyticlike effects of, albeit non-selective, D3R-preferring agonists (Rogers et al., 2000; Rogoz et al., 2003). Therefore, an alternative explanation for these changes is that they reflect a more general failure to inhibit behavior responses to novelty (see also Section 6.2). Indeed, indicating that D3R KO mice may harbor dysfunctions at the level of the prefrontal cortex, a brain region mediating in response inhibition, these mice exhibit working memory deficits and blunted prefrontal neuronal responses to D1R-like agonist treatment (Glickstein et al., 2002). In the rodent brain the D3R and D4R are expressed at much lower levels than the D2R (see also Section 6.1). In comparison to D2R, the D3R shows a highly restricted expression in limbic regions, including the nucleus accumbens (Sokoloff et al., 1990; Bouthenet et al., 1991; Levesque et al., 1992). This dense localization of D3R in the mescocortical circuit suggests that the receptor may play a prominent role in mediating reward-related behavior (for review, see Everitt et al., 1999). Indeed, drugs acting at the D3R are potent modulators of the reinforcing effects of psychostimulants and functional changes in D3R function are associated with altered responses to these drugs (Caine and Koob, 1993; Pilla et al., 1999; Xu et al., 2000a; Vorel et al., 2002). Further supporting a role for the D3R in reward, D3R KO mice show increased conditioned place preference to amphetamine and morphine (Xu et al., 1997; Narita et al., 2003). Moreover, suggesting D3R contribution to the reinforcing effects of cannabinoids, the ability of a cannabinoid CB1 receptor antagonist to activate nucleus accumbens neurons is diminished in D3R KO mice, but not D2R KOs (Duarte et al., 2003). Finally, one report has demonstrated reduced voluntary ethanol consumption, coupled with exaggerated ethanol-withdrawal and increased sensitivity to the hypnotic effects of ethanol in D3R KO mice (Narita et al., 2002). However, neither altered ethanol consumption, nor ethanol-induced conditioned place preference could be replicated in separate studies of D3R KO mice (Benoit et al., 2003; McQuade et al., 2003). Clearly, while additional research is needed, recent findings identify the D3R as an important component of the neural circuitry of reward and encourage further research in rewardrelated phenotypes in D3R KO mice.
1125
5.3. Supporting D1R-like/D2R-like cooperativity? D1R and D2R are present on the same striatal neurons, and while generally having mutually opposing effects on signal transduction, demonstrate cooperative/ synergistic effects under certain conditions (Walters et al., 1987; LaHoste et al., 1993, 2000 Xu et al., 1994a; Aizman et al., 2000). A supporting role for the D3R in modulating D1R/D2R cooperativity has been proposed (for discussion, see Glickstein and Schmauss, 2001). In this context, D3R KO mice show normal locomotor responses to treatment with either D1R-like or D2Rlike agents administered alone, but exaggerated hyperactivity responses to combined D1R-like/D2R-like agonist treatment (Accili et al., 1996; Xu et al., 1997; Betancur et al., 2001; Wong et al., 2003b; Yarkov et al., 2003). These findings are congruent with the idea that the D3R may normally inhibit the synergistic effects of D1R/D2R co-activation. However, a full picture of these processes is yet to emerge. For example, the synergistic hypoactive/cataleptic effects of D1R-likeþ D2R-like antagonist co-administration appear to be unaltered in D3R KO mice (Boulay et al., 2000). The ability of combined D1R-like þ D2R-like agonist administration to synergistically activate striatal neurons is also normal in D3R KO mice (Xu et al., 1997). Another approach to try and clarify the nature of functional interactions between the D1R, D2R and D3R has been to generate KO mice lacking the D3R and either the D1R or the D2R. Studies of D1Rþ D3R double KO mice are currently inconclusive, with reports showing that locomotor activity in double KOs is either no different from that seen in D1R KO mice, or that the hyperactivity observed in D1R KOs is further potentiated in mice lacking both D1R and D3R (Karasinska et al., 2000; Wong et al., 2003b). There is also some evidence that deletion of the D3R potentiates locomotor hypoactivity in D2R KO mice (Jung et al., 1999; but see Vallone et al., 2002). Elevations in striatal dopamine turnover and calbindin found in D2R KO mice have also been found to be augmented in D2R þ D3R double KO mice (Jung et al., 1999, 2000). These data suggest that the D2R and D3R normally mediate striatal function in a cooperative manner. This hypothesis would be consistent with the finding that striatal c-fos responses to D1R-like agonist stimulation are more effectively blocked by a combination of D3R gene deletion and D2R-antagonist treatment than either manipulation alone (Jung and Schmauss, 1999; Schmauss, 2000; Glickstein et al., 2002).
1126
A. Holmes et al. / Neuropharmacology 47 (2004) 1117–1134
6. D4R KO mice; losing (and gaining) their inhibitions 6.1. General In the only line of D4R KO mice currently available, mutant mice are viable, reproduce normally and show no gross morphological or neurological abnormalities (Rubinstein et al., 1997). Demonstrating an important role for the D4R in the regulation of adaptive retinal responses to changing levels of illumination, both basal and D2R-like agonist-induced photoreceptor responsivity is compromised in D4R KO mice (Nir et al., 2002). The D4R is expressed in the rodent striatum, albeit at significantly lower levels than the D2R (Van Tol et al., 1991; Primus et al., 1997; Khan et al., 1998). There is also a high density of D4R in rodent corticolimbic regions, including the prefrontal cortex, hippocampus, amygdala, and hypothalamus (MeadorWoodruff et al., 1996; Mrzljak et al., 1996; Ariano et al., 1997). Consistent with this relatively low abundance of the D4R, D2R-like binding is not demonstrably reduced in D4R KO mice (Rubinstein et al., 1997). However, D4R KO mice have been less extensively studied than other dopamine receptor KO mice and possible functional changes at other levels of the dopaminergic system, or in non-dopaminergic systems are yet to be thoroughly explored. A recent study has shown that both D1R and NMDA receptor binding is significantly increased in striatal regions on the D4R KO brain (Gan et al., 2004). Other work has demonstrated that a D4R-preferring agonist blocks the activity of GABAergic medium spiny neurons in the globus pallidus (via suppression of protein kinase A) and, conversely, that dopaminergic inhibition of these neurons is absent in D4R KO mice (Shin et al., 2003). This novel latter finding suggests that D4R may normally disinhibit basal ganglia output. 6.2. Behavior D4R KO mice show rotarod motor coordination superior to that of WT controls, exhibit exaggerated hyperactivity responses to both psychostimulants and ethanol, and are better at discriminating cocaine from saline than WTs (Rubinstein et al., 2001; Katz et al., 2003). One simplistic interpretation of these findings is that D4Rs may normally exert a generally inhibitory action on dopaminergic neurotransmission in certain brain regions. In agreement with this possibility, dopamine synthesis and turnover have been shown to be increased in the striatum, but not nucleus accumbens or frontal cortex, of D4R KO mice (Rubinstein et al., 1997, 2001). Moreover, electrophysiological studies
have demonstrated increased excitability of glutamatergic pyramidal neurons in the prefrontal cortex (but not striatum) of D4R KOs (Cepeda et al., 2001; Rubinstein et al., 2001). Finally, D4R KO mice exhibit increased susceptibility to the pro-convulsant effects of GABAA receptor antagonism (Rubinstein et al., 2001). This latter finding resembles the increased susceptibility to the epileptogenic and excitotoxic effects of glutamatergic or muscarinic receptor-challenge in D2R KO mice (Bozzi et al., 2000; Bozzi and Borrelli, 2002), and further supports a significant role for both the D2R and D4R in mediating neuronal excitability. While absence of the D4R may lead to a disinhibition of neural responses and increased sensitivity to stimulant drugs, research in both humans and D4R KO mice suggests a link between reduced D4R gene function and an abnormally high aversion to novelty. As noted above, across species, the D4R is densely expressed in the frontal cortex and limbic regions governing responses to novel and emotionally provocative stimuli (Meador-Woodruff et al., 1996; Mrzljak et al., 1996; Ariano et al., 1997). There is some evidence that individuals carrying a D4R gene variant that encodes for diminished ligand sensitivity show abnormally low levels of novelty-seeking (for review, see Paterson et al., 1999). In D4R KO mice, there are reports of moderately diminished open field locomotor activity, which may reflect an attenuated response to environmental novelty (Rubinstein et al., 1997; Katz et al., 2003). More saliently, specific investigation of novelty-seeking behavior in D4R KO mice has demonstrated heightened behavioral inhibition in high-novelty exploratory situations in mutant mice (Dulawa et al., 1999). Given that D4R KO mice were found to be normal in lownovelty, high-anxiety-provoking situations in this study, these phenotypic abnormalities appear to be specific to novelty, rather than anxiety (Dulawa et al., 1999). However, a subsequent study has found that D4R KO mice do display behavioral alterations consistent with increased anxiety-like behavior (Falzone et al., 2002). Therefore, the nature of changes in novelty-seeking, emotionality and behavioral inhibition in D4R KO mice is yet to be fully resolved.
7. Conclusions and future directions Phenotypic analysis of dopamine receptor KO mice has undoubtedly added to our understanding of how dopamine receptors function in the nervous system. In some cases this research has reinforced existing hypotheses regarding subtype function, for example, that the D2R is the prepotent autoreceptor controlling dopamine release and the D1R is integral to the behavioral and neural effects of psychostimulants. In other cases, dopamine receptor KO phenotypes appear to challenge
A. Holmes et al. / Neuropharmacology 47 (2004) 1117–1134
preexisting ideas about subtype function, largely based on data from imperfect pharmacological tools (e.g., the D2R, rather than the D3R, mediates dopamineinduced hypothermia). Findings in dopamine receptor KO mice have also nurtured new avenues of research. KO mice have provided uniquely valuable tools for dissecting the functional specificity of the D2RS and D2RL isoforms, and the distinct role of the D5R. The various successes advocate the continued use of dopamine receptor KO mice. Research on dopamine receptor KO mice has been at the forefront of the debate regarding the utility of the gene targeting approach to study the brain. Since the first dopamine receptor KO mice were reported on in 1994, limitations of this approach have become apparent. As discussed earlier, variability in behavioral phenotypic abnormalities observed across laboratories has been linked to the influences of genetic background and variations in testing procedure (see Sections 2, 4.3 and 4.4). Perhaps an even thornier issue is the reason for discrepancies between the effects of genetic vs. pharmacological inactivation of a dopamine receptor. There are a number of examples where phenotypic alterations in dopamine receptor KO mice differ from the predicted effects of pharmacological antagonists. These differences can be quantitative, as in the modest decrease in basal locomotor activity in D2R KO mice relative to the profound hypoactivity/catalepsyinducing effects of D2R-like antagonists in normal mice, and they can be qualitative, as in the hyperactive phenotype in D1R KO mice that is opposite to the hypoactivity/catalepsy-inducing effects of D1R-like antagonists. The fact that the locomotor effects of D2R-like and D1R-like antagonists are effectively abolished in D2R mice and D1R KO mice, respectively, makes it unlikely that they are normally mediated through other subtypes in each sub-family. Rather, these discrepancies suggest that adaptive changes have occurred in KO mice as a result of lifelong deletion of a receptor. In hindsight, the extent of these changes is perhaps not surprising given the significant plasticity of the brain, especially during development. Indeed, identifying the nature of compensatory processes could itself provide valuable insight into the plasticity and functional interconnectivity of the dopamine system. However, adaptive changes are clearly undesirable when the goal is to understand how a receptor subtype normally functions ceteris paribus. Because of the potential for the recruitment of compensatory changes to occlude or otherwise alter the normal function of a dopamine receptor subtype, it is now widely accepted that phenotypic abnormalities in KO mice must be interpreted with this caveat in mind. There are also a number of emerging approaches that can help mitigate the problem of compensation in dopamine receptor KO mice. One powerful approach is to restrict
1127
inactivation of dopamine receptor genes to specific brain regions or cell types. Induction of such ‘‘conditional’’ inactivation can also be withheld until after development to further limit the occurrence of adaptive changes (Heyer et al., 2002). Alternatives to gene targeting are also available, including antisense knockdown of dopamine receptors (Sibley, 1999) and, more recently, viral-mediated RNA interference of dopamine function (Hommel et al., 2003). Lastly, the development of novel highly selective dopamine receptor subtype agonists and antagonists will have a major impact on the field. Even with these important advances, researchers now recognize that KO mice are no research panacea. Dopamine receptor KO mice are most effectively utilized when employed in conjunction with other techniques, where available, to generate converging lines of evidence regarding dopamine receptor function. Understanding functional specification of dopamine receptor subtypes remains a fundamental question in neuroscience. Progress in this field of research has profound implications for the improved understanding and treatment of many common neurological and neuropsychiatric diseases. There is still much to accomplish to define the role of dopamine receptor subtypes in these disorders using KO mice. Of particular note, the role of specific dopamine receptor subtypes localized in the prefrontal cortex in mediating a variety of behaviors including cognition and reward is currently under-explored, as is the role of cortical and limbic dopamine receptors in the regulation of emotion and stress. These studies could contribute significantly to the development of novel pharmacotherapies for schizophrenia, addiction and mood disorders. Likewise, there are still significant advances to be made in understanding the nature of striatal dopamine dysfunction in Parkinson’s disease and other neurodegenerative movement disorders. We envision that dopamine receptor KO mice will continue to contribute a major component to research directed at meeting these objectives.
Acknowledgements DRS and AH are supported by the NIH Intramural Research Program.
References Accili, D., Fishburn, C.S., Drago, J., Steiner, H., Lachowicz, J.E., Park, B.H., Gauda, E.B., Lee, E.J., Cool, M.H., Sibley, D.R., Gerfen, C.R., Westphal, H., Fuchs, S., 1996. A targeted mutation of the D3 dopamine receptor gene is associated with hyperactivity in mice. Proceedings of the National Academy of Sciences USA 93, 1945–1949.
1128
A. Holmes et al. / Neuropharmacology 47 (2004) 1117–1134
Aizman, O., Brismar, H., Uhlen, P., Zettergren, E., Levey, A.I., Forssberg, H., Greengard, P., Aperia, A., 2000. Anatomical and physiological evidence for D1 and D2 dopamine receptor colocalization in neostriatal neurons. Nature Neuroscience 3, 226–230. Aoyama, S., Kase, H., Borrelli, E., 2000. Rescue of locomotor impairment in dopamine D2 receptor-deficient mice by an adenosine A2A receptor antagonist. The Journal of Neuroscience 20, 5848–5852. Apostolakis, E.M., Garai, J., Fox, C., Smith, C.L., Watson, S.J., Clark, J.H., O’Malley, B.W., 1996. Dopaminergic regulation of progesterone receptors: brain D5 dopamine receptors mediate induction of lordosis by D1-like agonists in rats. The Journal of Neuroscience 16, 4823–4834. Ariano, M.A., Sibley, D.R., 1994. Dopamine receptor distribution in the rat CNS: elucidation using anti-peptide antisera directed against D1A and D3 subtypes. Brain Research 649, 95–110. Ariano, M.A., Wang, J., Noblett, K.L., Larson, E.R., Sibley, D.R., 1997. Cellular distribution of the rat D4 dopamine receptor protein in the CNS using anti-receptor antisera. Brain Research 752, 26–34. Arnsten, A.F., Goldman-Rakic, P.S., 1998. Noise stress impairs prefrontal cortical cognitive function in monkeys: evidence for a hyperdopaminergic mechanism. Archives of General Psychiatry 55, 362–368. Asa, S.L., Kelly, M.A., Grandy, D.K., Low, M.J., 1999. Pituitary lactotroph adenomas develop after prolonged lactotroph hyperplasia in dopamine D2 receptor-deficient mice. Endocrinology 140, 5348–5355. Asico, L.D., Ladines, C., Fuchs, S., Accili, D., Carey, R.M., Semeraro, C., Pocchiari, F., Felder, R.A., Eisner, G.M., Jose, P.A., 1998. Disruption of the dopamine D3 receptor gene produces renin-dependent hypertension. Journal of Clinical Investigation 102, 493–498. Bach, M.E., Barad, M., Son, H., Zhuo, M., Lu, Y.F., Shih, R., Mansuy, I., Hawkins, R.D., Kandel, E.R., 1999. Age-related defects in spatial memory are correlated with defects in the late phase of hippocampal long-term potentiation in vitro and are attenuated by drugs that enhance the cAMP signaling pathway. Proceedings of the National Academy of Sciences USA 96, 5280–5285. Baik, J.H., Picetti, R., Saiardi, A., Thiriet, G., Dierich, A., Depaulis, A., Le Meur, M., Borrelli, E., 1995. Parkinsonian-like locomotor impairment in mice lacking dopamine D2 receptors. Nature 377, 424–428. Baufreton, J., Garret, M., Rivera, A., de la Calle, A., Gonon, F., Dufy, B., Bioulac, B., Taupignon, A., 2003. D5 (not D1) dopamine receptors potentiate burst-firing in neurons of the subthalamic nucleus by modulating an L-type calcium conductance. The Journal of Neuroscience 23, 816–825. Becker, A., Grecksch, G., Kraus, J., Peters, B., Schroeder, H., Schulz, S., Hollt, V., 2001. Loss of locomotor sensitisation in response to morphine in D1 receptor deficient mice. Naunyn Schmiedebergs Archives of Pharmacology 363, 562–568. Benoit-Marand, M., Borrelli, E., Gonon, F., 2001. Inhibition of dopamine release via presynaptic D2 receptors: time course and functional characteristics in vivo. The Journal of Neuroscience 21, 9134–9141. Benoit, S.C., McQuade, J.A., Clegg, D.J., Xu, M., Rushing, P.A., Woods, S.C., Seeley, R.J., 2003. Altered feeding responses in mice with targeted disruption of the dopamine-3 receptor gene. Behavioral Neuroscience 117, 46–54. Bergson, C., Mrzljak, L., Lidow, M.S., Goldman-Rakic, P.S., Levenson, R., 1995. Characterization of subtype-specific antibodies to the human D5 dopamine receptor: studies in primate brain and transfected mammalian cells. Proceedings of the National Academy of Sciences USA 92, 3468–3472. Betancur, C., Lepee-Lorgeoux, I., Cazillis, M., Accili, D., Fuchs, S., Rostene, W., 2001. Neurotensin gene expression and behavioral
responses following administration of psychostimulants and antipsychotic drugs in dopamine D(3) receptor deficient mice. Neuropsychopharmacology 24, 170–182. Boulay, D., Depoortere, R., Perrault, G., Borrelli, E., Sanger, D.J., 1999a. Dopamine D2 receptor knock-out mice are insensitive to the hypolocomotor and hypothermic effects of dopamine D2/D3 receptor agonists. Neuropharmacology 38, 1389–1396. Boulay, D., Depoortere, R., Rostene, W., Perrault, G., Sanger, D.J., 1999b. Dopamine D3 receptor agonists produce similar decreases in body temperature and locomotor activity in D3 knock-out and wild-type mice. Neuropharmacology 38, 555–565. Boulay, D., Depoortere, R., Oblin, A., Sanger, D.J., Schoemaker, H., Perrault, G., 2000. Haloperidol-induced catalepsy is absent in dopamine D(2), but maintained in dopamine D(3) receptor knock-out mice. European Journal of Pharmacology 391, 63–73. Bouthenet, M.L., Souil, E., Martres, M.P., Sokoloff, P., Giros, B., Schwartz, J.C., 1991. Localization of dopamine D3 receptor mRNA in the rat brain using in situ hybridization histochemistry: comparison with dopamine D2 receptor mRNA. Brain Research 564, 203–219. Boyce-Rustay, J.M., Risinger, F.O., 2003. Dopamine D3 receptor knockout mice and the motivational effects of ethanol. Pharmacology, Biochemistry and Behavior 75, 373–379. Bozzi, Y., Borrelli, E., 1999. Absence of the dopamine D2 receptor leads to a decreased expression of GDNF and NT-4 mRNAs in restricted brain areas. European Journal of Neuroscience 11, 1275–1284. Bozzi, Y., Borrelli, E., 2002. Dopamine D2 receptor signaling controls neuronal cell death induced by muscarinic and glutamatergic drugs. Molecular and Cellular Neuroscience 19, 263–271. Bozzi, Y., Vallone, D., Borrelli, E., 2000. Neuroprotective role of dopamine against hippocampal cell death. The Journal of Neuroscience 20, 8643–8649. Braff, D.L., Geyer, M.A., Swerdlow, N.R., 2001. Human studies of prepulse inhibition of startle: normal subjects, patient groups, and pharmacological studies. Psychopharmacology (Berlin) 156, 234–258. Bucan, M., Abel, T., 2002. The mouse: genetics meets behaviour. Nature Reviews Genetics 3, 114–123. Caine, S.B., Koob, G.F., 1993. Modulation of cocaine self-administration in the rat through D-3 dopamine receptors. Science 260, 1814–1816. Caine, S.B., Negus, S.S., Mello, N.K., Patel, S., Bristow, L., Kulagowski, J., Vallone, D., Saiardi, A., Borrelli, E., 2002. Role of dopamine D2-like receptors in cocaine self-administration: studies with D2 receptor mutant mice and novel D2 receptor antagonists. The Journal of Neuroscience 22, 2977–2988. Calabresi, P., Saiardi, A., Pisani, A., Baik, J.H., Centonze, D., Mercuri, N.B., Bernardi, G., Borrelli, E., 1997. Abnormal synaptic plasticity in the striatum of mice lacking dopamine D2 receptors. Journal of Neuroscience 17, 4536–4544. Cannon, C.M., Palmiter, R.D., 2003. Reward without dopamine. The Journal of Neuroscience 23, 10827–10831. Carta, A.R., Gerfen, C.R., Steiner, H., 2000. Cocaine effects on gene regulation in the striatum and behavior: increased sensitivity in D3 dopamine receptor-deficient mice. Neuroreport 11, 2395–2399. Centonze, D., Grande, C., Saulle, E., Martin, A.B., Gubellini, P., Pavon, N., Pisani, A., Bernardi, G., Moratalla, R., Calabresi, P., 2003a. Distinct roles of D1 and D5 dopamine receptors in motor activity and striatal synaptic plasticity. The Journal of Neuroscience 23, 8506–8512. Centonze, D., Grande, C., Usiello, A., Gubellini, P., Erbs, E., Martin, A.B., Pisani, A., Tognazzi, N., Bernardi, G., Moratalla, R., Borrelli, E., Calabresi, P., 2003b. Receptor subtypes involved in the presynaptic and postsynaptic actions of dopamine on striatal interneurons. The Journal of Neuroscience 23, 6245–6254. Cepeda, C., Hurst, R.S., Altemus, K.L., Flores-Hernandez, J., Calvert, C.R., Jokel, E.S., Grandy, D.K., Low, M.J., Rubinstein,
A. Holmes et al. / Neuropharmacology 47 (2004) 1117–1134 M.S., Ariano, M.A., Levine, M.S., 2001. Facilitated glutamatergic transmission in the striatum of D2 dopamine receptor-deficient mice. Journal of Neurophysiology 85, 659–670. Chausmer, A.L., Katz, J.L., 2001. The role of D2-like dopamine receptors in the locomotor stimulant effects of cocaine in mice. Psychopharmacology (Berlin) 155, 69–77. Chausmer, A.L., Elmer, G.I., Rubinstein, M., Low, M.J., Grandy, D.K., Katz, J.L., 2002. Cocaine-induced locomotor activity and cocaine discrimination in dopamine D2 receptor mutant mice. Psychopharmacology (Berlin) 163, 54–61. Chen, J.F., Moratalla, R., Impagnatiello, F., Grandy, D.K., Cuellar, B., Rubinstein, M., Beilstein, M.A., Hackett, E., Fink, J.S., Low, M.J., Ongini, E., Schwarzschild, M.A., 2001. The role of the D(2) dopamine receptor (D(2)R) in A(2A) adenosine receptor (A(2A)R)-mediated behavioral and cellular responses as revealed by A(2A) and D(2) receptor knockout mice. Proceedings of the National Academy of Sciences USA 98, 1970–1975. Chudasama, Y., Robbins, T.W., 2003. Dissociable contributions of the orbitofrontal and infralimbic cortex to pavlovian autoshaping and discrimination reversal learning: further evidence for the functional heterogeneity of the rodent frontal cortex. The Journal of Neuroscience 23, 8771–8780. Ciliax, B.J., Nash, N., Heilman, C., Sunahara, R., Hartney, A., Tiberi, M., Rye, D.B., Caron, M.G., Niznik, H.B., Levey, A.I., 2000. Dopamine D(5) receptor immunolocalization in rat and monkey brain. Synapse 37, 125–145. Civelli, O., Bunzow, J.R., Grandy, D.K., 1993. Molecular diversity of the dopamine receptors. Annual Review of Pharmacology and Toxicology 33, 281–307. Clifford, J.J., Tighe, O., Croke, D.T., Sibley, D.R., Drago, J., Waddington, J.L., 1998. Topographical evaluation of the phenotype of spontaneous behaviour in mice with targeted gene deletion of the D1A dopamine receptor: paradoxical elevation of grooming syntax. Neuropharmacology 37, 1595–1602. Clifford, J.J., Tighe, O., Croke, D.T., Kinsella, A., Sibley, D.R., Drago, J., Waddington, J.L., 1999. Conservation of behavioural topography to dopamine D1-like receptor agonists in mutant mice lacking the D1A receptor implicates a D1-like receptor not coupled to adenylyl cyclase. Neuroscience 93, 1483–1489. Clifford, J.J., Usiello, A., Vallone, D., Kinsella, A., Borrelli, E., Waddington, J.L., 2000. Topographical evaluation of behavioural phenotype in a line of mice with targeted gene deletion of the D2 dopamine receptor. Neuropharmacology 39, 382–390. Clifford, J.J., Kinsella, A., Tighe, O., Rubinstein, M., Grandy, D.K., Low, M.J., Croke, D.T., Waddington, J.L., 2001. Comparative, topographically-based evaluation of behavioural phenotype and specification of D(1)-like:D(2) interactions in a line of incipient congenic mice with D(2) dopamine receptor ‘knockout’. Neuropsychopharmacology 25, 527–536. Crabbe, J.C., Wahlsten, D., Dudek, B.C., 1999. Genetics of mouse behavior: interactions with laboratory environment. Science 284, 1670–1672. Crawford, C.A., Drago, J., Watson, J.B., Levine, M.S., 1997. Effects of repeated amphetamine treatment on the locomotor activity of the dopamine D1A-deficient mouse. Neuroreport 8, 2523–2527. Crawley, J.N., 2000. What’s Wrong with my Mouse? Behavioral Phenotyping of Transgenic and Knockout mice. Wiley-Liss, New York. Cromwell, H.C., Berridge, K.C., Drago, J., Levine, M.S., 1998. Action sequencing is impaired in D1A-deficient mutant mice. European Journal of Neuroscience 10, 2426–2432. Crusio, W.E., 1996. Gene-targeting studies: new methods, old problems. Trends in Neuroscience 19, 186–187, (discussion 188–189). Cunningham, C.L., Howard, M.A., Gill, S.J., Rubinstein, M., Low, M.J., Grandy, D.K., 2000. Ethanol-conditioned place preference is reduced in dopamine D2 receptor-deficient mice. Pharmacology, Biochemistry and Behavior 67, 693–699.
1129
Davila, V., Yan, Z., Craciun, L.C., Logothetis, D., Sulzer, D., 2003. D3 dopamine autoreceptors do not activate G-protein-gated inwardly rectifying potassium channel currents in substantia nigra dopamine neurons. The Journal of Neuroscience 23, 5693–5697. Depoortere, R., 1999. Combined use of mutant mice and classical pharmacology to characterise receptor functions: the case of dopamine D2 and D3 receptors. Psychopharmacology (Berlin) 147, 20–21. Diaz-Torga, G., Feierstein, C., Libertun, C., Gelman, D., Kelly, M.A., Low, M.J., Rubinstein, M., Becu-Villalobos, D., 2002. Disruption of the D2 dopamine receptor alters GH and IGF-I secretion and causes dwarfism in male mice. Endocrinology 143, 1270–1279. Dickinson, S.D., Sabeti, J., Larson, G.A., Giardina, K., Rubinstein, M., Kelly, M.A., Grandy, D.K., Low, M.J., Gerhardt, G.A., Zahniser, N.R., 1999. Dopamine D2 receptor-deficient mice exhibit decreased dopamine transporter function but no changes in dopamine release in dorsal striatum. Journal of Neurochemistry 72, 148–156. Dockstader, C.L., Rubinstein, M., Grandy, D.K., Low, M.J., van der Kooy, D., 2001. The D2 receptor is critical in mediating opiate motivation only in opiate-dependent and withdrawn mice. European Journal of Neuroscience 13, 995–1001. Dracheva, S., Haroutunian, V., 2001. Locomotor behavior of dopamine D1 receptor transgenic/D2 receptor deficient hybrid mice. Brain Research 905, 142–151. Dracheva, S., Xu, M., Kelley, K.A., Haroutunian, V., Holstein, G.R., Haun, S., Silverstein, J.H., Sealfon, S.C., 1999. Paradoxical locomotor behavior of dopamine D1 receptor transgenic mice. Experimental Neurology 157, 169–179. Drago, J., Gerfen, C.R., Lachowicz, J.E., Steiner, H., Hollon, T.R., Love, P.E., Ooi, G.T., Grinberg, A., Lee, E.J., Huang, S.P., et al., 1994. Altered striatal function in a mutant mouse lacking D1A dopamine receptors. Proceedings of the National Academy of Sciences USA 91, 12564–12568. Drago, J., Gerfen, C.R., Westphal, H., Steiner, H., 1996. D1 dopamine receptor-deficient mouse: cocaine-induced regulation of immediate-early gene and substance P expression in the striatum. Neuroscience 74, 813–823. Drago, F., Contarino, A., Busa, L., 1999. The expression of neuropeptide-induced excessive grooming behavior in dopamine D1 and D2 receptor-deficient mice. European Journal of Pharmacology 365, 125–131. Duarte, C., Alonso, R., Bichet, N., Cohen, C., Soubrie, P., Thiebot, M.H., 2004. Blockade by the cannabinoid CB1 receptor antagonist, Rimonabant (SR141716), of the potentiation by quinelorane of food-primed reinstatement of food-seeking behavior. Neuropsychopharmacology 29, 911–920. Dulawa, S.C., Grandy, D.K., Low, M.J., Paulus, M.P., Geyer, M.A., 1999. Dopamine D4 receptor-knock-out mice exhibit reduced exploration of novel stimuli. Journal of Neuroscience 19, 9550–9556. El-Ghundi, M., George, S.R., Drago, J., Fletcher, P.J., Fan, T., Nguyen, T., Liu, C., Sibley, D.R., Westphal, H., O’Dowd, B.F., 1998. Disruption of dopamine D1 receptor gene expression attenuates alcohol-seeking behavior. European Journal of Pharmacology 353, 149–158. El-Ghundi, M., Fletcher, P.J., Drago, J., Sibley, D.R., O’Dowd, B.F., George, S.R., 1999. Spatial learning deficit in dopamine D(1) receptor knockout mice. European Journal of Pharmacology 383, 95–106. El-Ghundi, M., O’Dowd, B.F., George, S.R., 2001. Prolonged fear responses in mice lacking dopamine D1 receptor. Brain Research 892, 86–93. El-Ghundi, M., O’Dowd, B.F., Erclik, M., George, S.R., 2003. Attenuation of sucrose reinforcement in dopamine D1 receptor deficient mice. European Journal of Neuroscience 17, 851–862.
1130
A. Holmes et al. / Neuropharmacology 47 (2004) 1117–1134
Elliot, E.E., Sibley, D.R., Katz, J.L., 2003. Locomotor and discriminative-stimulus effects of cocaine in dopamine D5 receptor knockout mice. Psychopharmacology (Berlin) 169, 161–168. Elmer, G.I., Pieper, J.O., Rubinstein, M., Low, M.J., Grandy, D.K., Wise, R.A., 2002. Failure of intravenous morphine to serve as an effective instrumental reinforcer in dopamine D2 receptor knockout mice. The Journal of Neuroscience 22, RC224. Everitt, B.J., Robbins, T.W., 1997. Central cholinergic systems and cognition. Annual Review of Psychology 48, 649–684. Everitt, B.J., Parkinson, J.A., Olmstead, M.C., Arroyo, M., Robledo, P., Robbins, T.W., 1999. Associative processes in addiction and reward. The role of amygdala-ventral striatal subsystems. Annals of the New York Academy of Sciences 877, 412–438. Falzone, T.L., Gelman, D.M., Young, J.I., Grandy, D.K., Low, M.J., Rubinstein, M., 2002. Absence of dopamine D4 receptors results in enhanced reactivity to unconditioned, but not conditioned, fear. European Journal of Neuroscience 15, 158–164. Fetsko, L.A., Xu, R., Wang, Y., 2003. Alterations in D1/D2 synergism may account for enhanced stereotypy and reduced climbing in mice lacking dopamine D2L receptor. Brain Research 967, 191–200. Friedman, E., Jin, L.Q., Cai, G.P., Hollon, T.R., Drago, J., Sibley, D.R., Wang, H.Y., 1997. D1-like dopaminergic activation of phosphoinositide hydrolysis is independent of D1A dopamine receptors: evidence from D1A knockout mice. Molecular Pharmacology 51, 6–11. Gan, L., Falzone, T.L., Zhang, K., Rubinstein, M., Ldessarini, R.J., Tarazi, F.I., 2004. Enhanced expression of dopamine D1 and glutamate NMDA receptors in dopamine D4 receptor knockout mice. Journal of Molecular Neuroscience 22, 167–178. Gerlai, R., 1996. Gene-targeting studies of mammalian behavior: is it the mutation or the background genotype? Trends in Neuroscience 19, 177–181. Geyer, M.A., McIlwain, K.L., Paylor, R., 2002. Mouse genetic models for prepulse inhibition: an early review. Molecular Psychiatry 7, 1039–1053. Gingrich, J.A., Caron, M.G., 1993. Recent advances in the molecular biology of dopamine receptors. Annual Review of Neuroscience 16, 299–321. Giros, B., Sokoloff, P., Martres, M.P., Riou, J.F., Emorine, L.J., Schwartz, J.C., 1989. Alternative splicing directs the expression of two D2 dopamine receptor isoforms. Nature 342, 923–926. Glickstein, S.B., Schmauss, C., 2001. Dopamine receptor functions: lessons from knockout mice [corrected]. Pharmacology and Therapeutics 91, 63–83. Glickstein, S.B., Schmauss, C., 2004. Focused motor stereotypies do not require enhanced activation of neurons in striosomes. Journal of Comparative Neurology 469, 227–238. Glickstein, S.B., Hof, P.R., Schmauss, C., 2002. Mice lacking dopamine D2 and D3 receptors have spatial working memory deficits. The Journal of Neuroscience 22, 5619–5629. Goldman-Rakic, P.S., 1995. Anatomical and functional circuits in prefrontal cortex of nonhuman primates. Relevance to epilepsy. Advances in Neurology 66, 51–63. Granon, S., Passetti, F., Thomas, K.L., Dalley, J.W., Everitt, B.J., Robbins, T.W., 2000. Enhanced and impaired attentional performance after infusion of D1 dopaminergic receptor agents into rat prefrontal cortex. The Journal of Neuroscience 20, 1208–1215. Gurden, H., Takita, M., Jay, T.M., 2000. Essential role of D1 but not D2 receptors in the NMDA receptor-dependent long-term potentiation at hippocampal-prefrontal cortex synapses in vivo. The Journal of Neuroscience 20, RC106. Hersi, A.I., Kitaichi, K., Srivastava, L.K., Gaudreau, P., Quirion, R., 2000. Dopamine D-5 receptor modulates hippocampal acetylcholine release. Brain Research Molecular Brain Research 76, 336–340.
Heyer, J., Xiao, Q., Bugaj-Gaweda, B., Ramboz, S., Unterbeck, A., 2002. Conditional inactivation of the dopamine receptor 5 gene: flanking the Drd5 gene with loxP sites. Genesis 32, 102–104. Hollon, T.R., Bek, M.J., Lachowicz, J.E., Ariano, M.A., Mezey, E., Ramachandran, R., Wersinger, S.R., Soares-da-Silva, P., Liu, Z.F., Grinberg, A., Drago, J., Young III, W.S., Westphal, H., Jose, P.A., Sibley, D.R., 2002. Mice lacking D5 dopamine receptors have increased sympathetic tone and are hypertensive. The Journal of Neuroscience 22, 10801–10810. Holmes, A., 2001. Targeted gene mutation approaches to the study of anxiety-like behavior in mice. Neuroscience and Biobehavioral Reviews 25, 261–273. Holmes, A., Rodgers, R.J., 1998. Responses of Swiss-Webster mice to repeated plus-maze experience: further evidence for a qualitative shift in emotional state? Pharmacology, Biochemistry and Behavior 60, 473–488. Holmes, A., Hollon, T.R., Gleason, T.C., Liu, Z., Dreiling, J., Sibley, D.R., Crawley, J.N., 2001. Behavioral characterization of dopamine D5 receptor null mutant mice. Behavioral Neuroscience 115, 1129–1144. Holmes, A., Wrenn, C.C., Harris, A.P., Thayer, K.E., Crawley, J.N., 2002. Behavioral profiles of inbred strains on novel olfactory, spatial and emotional tests for reference memory in mice. Genes, Brain and Behavior 1, 55–69. Holmes, A., Lit, Q., Murphy, D.L., Gold, E., Crawley, J.N., 2003. Abnormal anxiety-related behavior in serotonin transporter null mutant mice: the influence of genetic background. Genes, Brain and Behavior 2, 365–380. Hommel, J.D., Sears, R.M., Georgescu, D., Simmons, D.L., DiLeone, R.J., 2003. Local gene knockdown in the brain using viral-mediated RNA interference. Nature Medicine 9, 1539–1544. Huang, Y.Y., Simpson, E., Kellendonk, C., Kandel, E.R., 2004. Genetic evidence for the bidirectional modulation of synaptic plasticity in the prefrontal cortex by D1 receptors. Proceedings of the National Academy of Sciences USA 101, 3236–3241. Hunyady, B., Palkovits, M., Mozsik, G., Molnar, J., Feher, K., Toth, Z., Zolyomi, A., Szalayova, I., Key, S., Sibley, D.R., Mezey, E., 2001. Susceptibility of dopamine D5 receptor targeted mice to cysteamine. Journal of Physiology (Paris) 95, 147–151. Jose, P.A., Eisner, G.M., Felder, R.A., 1998. Regulation of D1 receptor function in spontaneous hypertension. Advances in Pharmacology 42, 525–528. Jung, M.Y., Schmauss, C., 1999. Decreased c-fos responses to dopamine D(1) receptor agonist stimulation in mice deficient for D(3) receptors. Journal of Biological Chemistry 274, 29406–29412. Jung, M.Y., Skryabin, B.V., Arai, M., Abbondanzo, S., Fu, D., Brosius, J., Robakis, N.K., Polites, H.G., Pintar, J.E., Schmauss, C., 1999. Potentiation of the D2 mutant motor phenotype in mice lacking dopamine D2 and D3 receptors. Neuroscience 91, 911–924. Jung, M.Y., Hof, P.R., Schmauss, C., 2000. Targeted disruption of the dopamine D(2) and D(3) receptor genes leads to different alterations in the expression of striatal calbindin-D(28k). Neuroscience 97, 495–504. Kalivas, P.W., Stewart, J., 1991. Dopamine transmission in the initiation and expression of drug- and stress-induced sensitization of motor activity. Brain Research Brain Research Reviews 16, 223–244. Karasinska, J.M., George, S.R., El-Ghundi, M., Fletcher, P.J., O’Dowd, B.F., 2000. Modification of dopamine D(1) receptor knockout phenotype in mice lacking both dopamine D(1) and D(3) receptors. European Journal of Pharmacology 399, 171–181. Karper, P.E., De la Rosa, H., Newman, E.R., Krall, C.M., Nazarian, A., McDougall, S.A., Crawford, C.A., 2002. Role of D1-like receptors in amphetamine-induced behavioral sensitization: a study using D1A receptor knockout mice. Psychopharmacology (Berlin) 159, 407–414.
A. Holmes et al. / Neuropharmacology 47 (2004) 1117–1134 Kase, H., Aoyama, S., Ichimura, M., Ikeda, K., Ishii, A., Kanda, T., Koga, K., Koike, N., Kurokawa, M., Kuwana, Y., Mori, A., Nakamura, J., Nonaka, H., Ochi, M., Saki, M., Shimada, J., Shindou, T., Shiozaki, S., Suzuki, F., Takeda, M., Yanagawa, K., Richardson, P.J., Jenner, P., Bedard, P., Borrelli, E., Hauser, R.A., Chase, T.N., 2003. Progress in pursuit of therapeutic A2A antagonists: the adenosine A2A receptor selective antagonist KW6002: research and development toward a novel nondopaminergic therapy for Parkinson’s disease. Neurology 61, S97–S100. Katz, J.L., Chausmer, A.L., Elmer, G.I., Rubinstein, M., Low, M.J., Grandy, D.K., 2003. Cocaine-induced locomotor activity and cocaine discrimination in dopamine D4 receptor mutant mice. Psychopharmacology (Berlin) 170, 108–114. Kebabian, J.W., Calne, D.B., 1979. Multiple receptors for dopamine. Nature 277, 93–96. Kelly, M.A., Rubinstein, M., Asa, S.L., Zhang, G., Saez, C., Bunzow, J.R., Allen, R.G., Hnasko, R., Ben-Jonathan, N., Grandy, D.K., Low, M.J., 1997. Pituitary lactotroph hyperplasia and chronic hyperprolactinemia in dopamine D2 receptor-deficient mice. Neuron 19, 103–113. Kelly, M.A., Rubinstein, M., Phillips, T.J., Lessov, C.N., BurkhartKasch, S., Zhang, G., Bunzow, J.R., Fang, Y., Gerhardt, G.A., Grandy, D.K., Low, M.J., 1998. Locomotor activity in D2 dopamine receptor-deficient mice is determined by gene dosage, genetic background, and developmental adaptations. The Journal of Neuroscience 18, 3470–3479. Khan, Z.U., Gutierrez, A., Martin, R., Penafiel, A., Rivera, A., De La Calle, A., 1998. Differential regional and cellular distribution of dopamine D2-like receptors: an immunocytochemical study of subtype-specific antibodies in rat and human brain. Journal of Comparative Neurology 402, 353–371. Koeltzow, T.E., Xu, M., Cooper, D.C., Hu, X.T., Tonegawa, S., Wolf, M.E., White, F.J., 1998. Alterations in dopamine release but not dopamine autoreceptor function in dopamine D3 receptor mutant mice. The Journal of Neuroscience 18, 2231–2238. LaHoste, G.J., Yu, J., Marshall, J.F., 1993. Striatal Fos expression is indicative of dopamine D1/D2 synergism and receptor supersensitivity. Proceedings of the National Academy of Sciences USA 90, 7451–7455. LaHoste, G.J., Henry, B.L., Marshall, J.F., 2000. Dopamine D1 receptors synergize with D2, but not D3 or D4, receptors in the striatum without the involvement of action potentials. The Journal of Neuroscience 20, 6666–6671. Laplante, F., Sibley, D.R., Quirion, R., 2004. Reduction in acetylcholine release in the hippocampus of dopamine D5 receptor deficient mice. Neuropsychopharmacology, 29, 1620–1627. Levesque, D., Diaz, J., Pilon, C., Martres, M.P., Giros, B., Souil, E., Schott, D., Morgat, J.L., Schwartz, J.C., Sokoloff, P., 1992. Identification, characterization, and localization of the dopamine D3 receptor in rat brain using 7-[3H]hydroxy-N,N-di-n-propyl-2aminotetralin. Proceedings of the National Academy of Sciences USA 89, 8155–8159. Levey, A.I., Hersch, S.M., Rye, D.B., Sunahara, R.K., Niznik, H.B., Kitt, C.A., Price, D.L., Maggio, R., Brann, M.R., Ciliax, B.J., et al., 1993. Localization of D1 and D2 dopamine receptors in brain with subtype-specific antibodies. Proceedings of the National Academy of Sciences USA 90, 8861–8865. L’Hirondel, M., Cheramy, A., Godeheu, G., Artaud, F., Saiardi, A., Borrelli, E., Glowinski, J., 1998. Lack of autoreceptor-mediated inhibitory control of dopamine release in striatal synaptosomes of D2 receptor-deficient mice. Brain Research 792, 253–262. Li, X.X., Bek, M., Asico, L.D., Yang, Z., Grandy, D.K., Goldstein, D.S., Rubinstein, M., Eisner, G.M., Jose, P.A., 2001. Adrenergic and endothelin B receptor-dependent hypertension in dopamine receptor type-2 knockout mice. Hypertension 38, 303–308.
1131
Li, X., Bek, M., Asico, L.D., Yang, Z., Rubinstein, M., Grandy, D.K., Goldstein, D.S., Eisner, G.M., Jose, P.A., 2002. [Disruption of the D(2) dopamine receptor produces adrenergic and endothelin B receptor-dependent hypertension]. Zhonghua Yi Xue Za Zhi 82, 703–707. Li, S., Cullen, W.K., Anwyl, R., Rowan, M.J., 2003. Dopaminedependent facilitation of LTP induction in hippocampal CA1 by exposure to spatial novelty. Nature Neuroscience 6, 526–531. Lidow, M.S., Koh, P.O., Arnsten, A.F., 2003. D1 dopamine receptors in the mouse prefrontal cortex: immunocytochemical and cognitive neuropharmacological analyses. Synapse 47, 101–108. Lindgren, N., Usiello, A., Goiny, M., Haycock, J., Erbs, E., Greengard, P., Hokfelt, T., Borrelli, E., Fisone, G., 2003. Distinct roles of dopamine D2L and D2S receptor isoforms in the regulation of protein phosphorylation at presynaptic and postsynaptic sites. Proceedings of the National Academy of Sciences USA 100, 4305–4309. Lindvall, O., Bjorklund, A., 1978. Anatomy of the dopaminergic neuron systems in the rat brain. Advances in Biochemistry and Psychopharmacology 19, 1–23. Mack, K.J., O’Malley, K.L., Todd, R.D., 1991. Differential expression of dopaminergic D2 receptor messenger RNAs during development. Brain Research Developmental Brain Research 59, 249–251. MacQueen, G.M., Campbell, S., McEwen, B.S., Macdonald, K., Amano, S., Joffe, R.T., Nahmias, C., Young, L.T., 2003. Course of illness, hippocampal function, and hippocampal volume in major depression. Proceedings of the National Academy of Sciences USA 100, 1387–1392. Maldonado, R., Saiardi, A., Valverde, O., Samad, T.A., Roques, B.P., Borrelli, E., 1997. Absence of opiate rewarding effects in mice lacking dopamine D2 receptors. Nature 388, 586–589. Matthies, H., Becker, A., Schroeder, H., Kraus, J., Hollt, V., Krug, M., 1997. Dopamine D1-deficient mutant mice do not express the late phase of hippocampal long-term potentiation. Neuroreport 8, 3533–3535. McIlwain, K.L., Merriweather, M.Y., Yuva-Paylor, L.A., Paylor, R., 2001. The use of behavioral test batteries: effects of training history. Physiological Behavior 73, 705–717. McNamara, F.N., Clifford, J.J., Tighe, O., Kinsella, A., Drago, J., Fuchs, S., Croke, D.T., Waddington, J.L., 2002. Phenotypic, ethologically based resolution of spontaneous and D(2)-like vs D(1)-like agonist-induced behavioural topography in mice with congenic D(3) dopamine receptor ‘‘knockout’’. Synapse 46, 19–31. McNamara, F.N., Clifford, J.J., Tighe, O., Kinsella, A., Drago, J., Croke, D.T., Waddington, J.L., 2003. Congenic D1A dopamine receptor mutants: ethologically based resolution of behavioural topography indicates genetic background as a determinant of knockout phenotype. Neuropsychopharmacology 28, 86–99. McQuade, J.A., Xu, M., Woods, S.C., Seeley, R.J., Benoit, S.C., 2003. Ethanol consumption in mice with a targeted disruption of the dopamine-3 receptor gene. Addiction Biology 8, 295–303. McQuade, J.A., Benoit, S.C., Xu, M., Woods, S.C., Seeley, R.J., 2004. High-fat diet induced adiposity in mice with targeted disruption of the dopamine-3 receptor gene. Behavioral Brain Research 151, 313–319. Meador-Woodruff, J.H., Mansour, A., Bunzow, J.R., Van Tol, H.H., Watson Jr., S.J., Civelli, O., 1989. Distribution of D2 dopamine receptor mRNA in rat brain. Proceedings of the National Academy of Sciences USA 86, 7625–7628. Meador-Woodruff, J.H., Damask, S.P., Wang, J., Haroutunian, V., Davis, K.L., Watson, S.J., 1996. Dopamine receptor mRNA expression in human striatum and neocortex. Neuropsychopharmacology 15, 17–29. Mercuri, N.B., Saiardi, A., Bonci, A., Picetti, R., Calabresi, P., Bernardi, G., Borrelli, E., 1997. Loss of autoreceptor function in
1132
A. Holmes et al. / Neuropharmacology 47 (2004) 1117–1134
dopaminergic neurons from dopamine D2 receptor deficient mice. Neuroscience 79, 323–327. Milad, M.R., Quirk, G.J., 2002. Neurons in medial prefrontal cortex signal memory for fear extinction. Nature 420, 70–74. Millan, M.J., Peglion, J.L., Vian, J., Rivet, J.M., Brocco, M., Gobert, A., Newman-Tancredi, A., Dacquet, C., Bervoets, K., Girardon, S., et al., 1995. Functional correlates of dopamine D3 receptor activation in the rat in vivo and their modulation by the selective antagonist, (+)-S 14297: 1. Activation of postsynaptic D3 receptors mediates hypothermia, whereas blockade of D2 receptors elicits prolactin secretion and catalepsy. The Journal of Pharmacology and Experimental Therapeutics 275, 885–898. Miner, L.L., Drago, J., Chamberlain, P.M., Donovan, D., Uhl, G.R., 1995. Retained cocaine conditioned place preference in D1 receptor deficient mice. Neuroreport 6, 2314–2316. Missale, C., Nash, S.R., Robinson, S.W., Jaber, M., Caron, M.G., 1998. Dopamine receptors: from structure to function. Physiological Reviews 78, 189–225. Miyamoto, S., Mailman, R.B., Lieberman, J.A., Duncan, G.E., 2001. Blunted brain metabolic response to ketamine in mice lacking D(1A) dopamine receptors. Brain Research 894, 167–180. Mizuo, K., Narita, M., Miyatake, M., Suzuki, T., 2004. Enhancement of dopamine-induced signaling responses in the forebrain of mice lacking dopamine D3 receptor. Neuroscience Letters 358, 13–16. Monsma Jr., F.J., McVittie, L.D., Gerfen, C.R., Mahan, L.C., Sibley, D.R., 1989. Multiple D2 dopamine receptors produced by alternative RNA splicing. Nature 342, 926–929. Montague, D.M., Striplin, C.D., Overcash, J.S., Drago, J., Lawler, C.P., Mailman, R.B., 2001. Quantification of D1B(D5) receptors in dopamine D1A receptor-deficient mice. Synapse 39, 319–322. Moratalla, R., Xu, M., Tonegawa, S., Graybiel, A.M., 1996. Cellular responses to psychomotor stimulant and neuroleptic drugs are abnormal in mice lacking the D1 dopamine receptor. Proceedings of the National Academy of Sciences USA 93, 14928–14933. Mrzljak, L., Bergson, C., Pappy, M., Huff, R., Levenson, R., Goldman-Rakic, P.S., 1996. Localization of dopamine D4 receptors in GABAergic neurons of the primate brain. Nature 381, 245–248. Murer, M.G., Dziewczapolski, G., Salin, P., Vila, M., Tseng, K.Y., Ruberg, M., Rubinstein, M., Kelly, M.A., Grandy, D.K., Low, M.J., Hirsch, E., Raisman-Vozari, R., Gershanik, O., 2000. The indirect basal ganglia pathway in dopamine D(2) receptordeficient mice. Neuroscience 99, 643–650. Narita, M., Soma, M., Tamaki, H., Suzuki, T., 2002. Intensification of the development of ethanol dependence in mice lacking dopamine D(3) receptor. Neuroscience Letters 324, 129–132. Narita, M., Mizuo, K., Mizoguchi, H., Sakata, M., Tseng, L.F., Suzuki, T., 2003. Molecular evidence for the functional role of dopamine D3 receptor in the morphine-induced rewarding effect and hyperlocomotion. The Journal of Neuroscience 23, 1006– 1012. Neve, K.A., Neve, R.L., Fidel, S., Janowsky, A., Higgins, G.A., 1991. Increased abundance of alternatively spliced forms of D2 dopamine receptor mRNA after denervation. Proceedings of the National Academy of Sciences USA 88, 2802–2806. Nir, I., Harrison, J.M., Haque, R., Low, M.J., Grandy, D.K., Rubinstein, M., Iuvone, P.M., 2002. Dysfunctional light-evoked regulation of cAMP in photoreceptors and abnormal retinal adaptation in mice lacking dopamine D4 receptors. The Journal of Neuroscience 22, 2063–2073. Otani, S., Blond, O., Desce, J.M., Crepel, F., 1998. Dopamine facilitates long-term depression of glutamatergic transmission in rat prefrontal cortex. Neuroscience 85, 669–676. Ozono, R., Ueda, A., Oishi, Y., Yano, A., Kambe, M., Katsuki, M., Oshima, T., 2003. Dopamine D2 receptor modulates sodium handling via local production of dopamine in the kidney. Journal of Cardiovascular Pharmacology 42, S75–S79.
Palmer, A.A., Low, M.J., Grandy, D.K., Phillips, T.J., 2003. Effects of a Drd2 deletion mutation on ethanol-induced locomotor stimulation and sensitization suggest a role for epistasis. Behavior Genetics 33, 311–324. Parish, C.L., Stanic, D., Drago, J., Borrelli, E., Finkelstein, D.I., Horne, M.K., 2002. Effects of long-term treatment with dopamine receptor agonists and antagonists on terminal arbor size. European Journal of Neuroscience 16, 787–794. Paterson, A.D., Sunohara, G.A., Kennedy, J.L., 1999. Dopamine D4 receptor gene: novelty or nonsense? Neuropsychopharmacology 21, 3–16. Perachon, S., Betancur, C., Pilon, C., Rostene, W., Schwartz, J.C., Sokoloff, P., 2000. Role of dopamine D3 receptors in thermoregulation: a reappraisal. Neuroreport 11, 221–225. Phillips, T.J., Crabbe, J.C., Metten, P., Belknap, J.K., 1994. Localization of genes affecting alcohol drinking in mice. Alcoholism: Clinical and Experimental Research 18, 931–941. Phillips, T.J., Brown, K.J., Burkhart-Kasch, S., Wenger, C.D., Kelly, M.A., Rubinstein, M., Grandy, D.K., Low, M.J., 1998. Alcohol preference and sensitivity are markedly reduced in mice lacking dopamine D2 receptors. Nature Neuroscience 1, 610–615. Pilla, M., Perachon, S., Sautel, F., Garrido, F., Mann, A., Wermuth, C.G., Schwartz, J.C., Everitt, B.J., Sokoloff, P., 1999. Selective inhibition of cocaine-seeking behaviour by a partial dopamine D3 receptor agonist. Nature 400, 371–375. Primus, R.J., Thurkauf, A., Xu, J., Yevich, E., McInerney, S., Shaw, K., Tallman, J.F., Gallagher, D.W., 1997. II. Localization and characterization of dopamine D4 binding sites in rat and human brain by use of the novel, D4 receptor-selective ligand [3H]NGD 94-1. The Journal of Pharmacology and Experimental Therapeutics 282, 1020–1027. Pritchard, L.M., Logue, A.D., Hayes, S., Welge, J.A., Xu, M., Zhang, J., Berger, S.P., Richtand, N.M., 2003. 7-OH-DPAT and PD 128907 selectively activate the D3 dopamine receptor in a novel environment. Neuropsychopharmacology 28, 100–107. Ralph, R.J., Varty, G.B., Kelly, M.A., Wang, Y.M., Caron, M.G., Rubinstein, M., Grandy, D.K., Low, M.J., Geyer, M.A., 1999. The dopamine D2, but not D3 or D4, receptor subtype is essential for the disruption of prepulse inhibition produced by amphetamine in mice. The Journal of Neuroscience 19, 4627–4633. Ralph-Williams, R.J., Lehmann-Masten, V., Otero-Corchon, V., Low, M.J., Geyer, M.A., 2002. Differential effects of direct and indirect dopamine agonists on prepulse inhibition: a study in D1 and D2 receptor knock-out mice. The Journal of Neuroscience 22, 9604–9611. Risinger, F.O., Freeman, P.A., Rubinstein, M., Low, M.J., Grandy, D.K., 2000. Lack of operant ethanol self-administration in dopamine D2 receptor knockout mice. Psychopharmacology (Berlin) 152, 343–350. Robbins, T.W., 2000. Chemical neuromodulation of frontal-executive functions in humans and other animals. Experimental Brain Research 133, 130–138. Rodgers, R.J., Cao, B.J., Dalvi, A., Holmes, A., 1997. Animal models of anxiety: an ethological perspective. Brazilian Journal of Medical and Biological Research 30, 289–304. Rogers, D.C., Costall, B., Domeney, A.M., Gerrard, P.A., Greener, M., Kelly, M.E., Hagan, J.J., Hunter, A.J., 2000. Anxiolytic profile of ropinirole in the rat, mouse and common marmoset. Psychopharmacology (Berlin) 151, 91–97. Rogoz, Z., Skuza, G., Klodzinska, A., 2003. Anxiolytic-like effects of preferential dopamine D3 receptor agonists in an animal model. Polish Journal of Pharmacology 55, 449–454. Romach, M.K., Glue, P., Kampman, K., Kaplan, H.L., Somer, G.R., Poole, S., Clarke, L., Coffin, V., Cornish, J., O’Brien, C.P., Sellers, E.M., 1999. Attenuation of the euphoric effects of cocaine by the dopamine D1/D5 antagonist ecopipam (SCH 39166). Archives of General Psychiatry 56, 1101–1106.
A. Holmes et al. / Neuropharmacology 47 (2004) 1117–1134 Roth, R.H., 1984. CNS dopamine autoreceptors: distribution, pharmacology, and function. Annals of the New York Academy of Sciences 430, 27–53. Rouge-Pont, F., Usiello, A., Benoit-Marand, M., Gonon, F., Piazza, P.V., Borrelli, E., 2002. Changes in extracellular dopamine induced by morphine and cocaine: crucial control by D2 receptors. The Journal of Neuroscience 22, 3293–3301. Rubinstein, M., Phillips, T.J., Bunzow, J.R., Falzone, T.L., Dziewczapolski, G., Zhang, G., Fang, Y., Larson, J.L., McDougall, J.A., Chester, J.A., Saez, C., Pugsley, T.A., Gershanik, O., Low, M.J., Grandy, D.K., 1997. Mice lacking dopamine D4 receptors are supersensitive to ethanol, cocaine, and methamphetamine. Cell 90, 991–1001. Rubinstein, M., Cepeda, C., Hurst, R.S., Flores-Hernandez, J., Ariano, M.A., Falzone, T.L., Kozell, L.B., Meshul, C.K., Bunzow, J.R., Low, M.J., Levine, M.S., Grandy, D.K., 2001. Dopamine D4 receptor-deficient mice display cortical hyperexcitability. The Journal of Neuroscience 21, 3756–3763. Saiardi, A., Borrelli, E., 1998. Absence of dopaminergic control on melanotrophs leads to Cushing’s-like syndrome in mice. Molecular Endocrinology 12, 1133–1139. Saiardi, A., Bozzi, Y., Baik, J.H., Borrelli, E., 1997. Antiproliferative role of dopamine: loss of D2 receptors causes hormonal dysfunction and pituitary hyperplasia. Neuron 19, 115–126. Schmauss, C., 2000. A single dose of methamphetamine leads to a long term reversal of the blunted dopamine D1 receptor-mediated neocortical c-fos responses in mice deficient for D2 and D3 receptors. Journal of Biological Chemistry 275, 38944–38948. Schmauss, C., Glickstein, S.B., Adlersberg, M., Hsiung, S.C., Tamir, H., 2002. A single dose of methamphetamine rescues the blunted dopamine D(1)-receptor activity in the neocortex of D(2)- and D(3)-receptor knockout mice. Annals of the New York Academy of Sciences 965, 21–27. Schmitz, Y., Lee, C.J., Schmauss, C., Gonon, F., Sulzer, D., 2001. Amphetamine distorts stimulation-dependent dopamine overflow: effects on D2 autoreceptors, transporters, and synaptic vesicle stores. The Journal of Neuroscience 21, 5916–5924. Schmitz, Y., Schmauss, C., Sulzer, D., 2002. Altered dopamine release and uptake kinetics in mice lacking D2 receptors. The Journal of Neuroscience 22, 8002–8009. Schwartz, J.C., Levesque, D., Martres, M.P., Sokoloff, P., 1993. Dopamine D3 receptor: basic and clinical aspects. Clinical Neuropharmacology 16, 295–314. Shin, R.M., Masuda, M., Miura, M., Sano, H., Shirasawa, T., Song, W.J., Kobayashi, K., Aosaki, T., 2003. Dopamine D4 receptorinduced postsynaptic inhibition of GABAergic currents in mouse globus pallidus neurons. The Journal of Neuroscience 23, 11662–11672. Sibley, D.R., 1999. New insights into dopaminergic receptor function using antisense and genetically altered animals. Annual Review of Pharmacology and Toxicology 39, 313–341. Sibley, D.R., Creese, I., 1983. Regulation of ligand binding to pituitary D-2 dopaminergic receptors. Effects of divalent cations and functional group modification. Journal of Biological Chemistry 258, 4957–4965. Sibley, D.R., Monsma, Jr.F.J., Shen, Y., 1993. Molecular neurobiology of dopaminergic receptors. International Review of Neurobiology 35, 391–415. Smith, D.R., Striplin, C.D., Geller, A.M., Mailman, R.B., Drago, J., Lawler, C.P., Gallagher, M., 1998. Behavioural assessment of mice lacking D1A dopamine receptors. Neuroscience 86, 135–146. Smith, J.W., Fetsko, L.A., Xu, R., Wang, Y., 2002. Dopamine D2L receptor knockout mice display deficits in positive and negative reinforcing properties of morphine and in avoidance learning. Neuroscience 113, 755–765. Sokoloff, P., Giros, B., Martres, M.P., Bouthenet, M.L., Schwartz, J.C., 1990. Molecular cloning and characterization of a novel
1133
dopamine receptor (D3) as a target for neuroleptics. Nature 347, 146–151. Steiner, H., Fuchs, S., Accili, D., 1997. D3 dopamine receptordeficient mouse: evidence for reduced anxiety. Physiology and Behavior 63, 137–141. Svenningsson, P., Tzavara, E.T., Carruthers, R., Rachleff, I., Wattler, S., Nehls, M., McKinzie, D.L., Fienberg, A.A., Nomikos, G.G., Greengard, P., 2003. Diverse psychotomimetics act through a common signaling pathway. Science 302, 1412–1415. Tan, S., Hermann, B., Borrelli, E., 2003. Dopaminergic mouse mutants: investigating the roles of the different dopamine receptor subtypes and the dopamine transporter. International Review of Neurobiology 54, 145–197. Tarantino, L.M., Gould, T.J., Druhan, J.P., Bucan, M., 2000. Behavior and mutagenesis screens: the importance of baseline analysis of inbred strains. Mammalian Genome 11, 555–564. Tomiyama, K., McNamara, F.N., Clifford, J.J., Kinsella, A., Drago, J., Tighe, O., Croke, D.T., Koshikawa, N., Waddington, J.L., 2002. Phenotypic resolution of spontaneous and D1-like agonist-induced orofacial movement topographies in congenic dopamine D1A receptor ‘knockout’ mice. Neuropharmacology 42, 644–652. Tomiyama, K., McNamara, F.N., Clifford, J.J., Kinsella, A., Drago, J., Fuchs, S., Grandy, D.K., Low, M.J., Rubinstein, M., Tighe, O., Croke, D.T., Koshikawa, N., Waddington, J.L., 2004. Comparative phenotypic resolution of spontaneous, D2-like and D1like agonist-induced orofacial movement topographies in congenic mutants with dopamine D2 vs. D3 receptor ‘‘knockout’’. Synapse 51, 71–81. Tran, A.H., Tamura, R., Uwano, T., Kobayashi, T., Katsuki, M., Matsumoto, G., Ono, T., 2002. Altered accumbens neural response to prediction of reward associated with place in dopamine D2 receptor knockout mice. Proceedings of the National Academy of Sciences USA 99, 8986–8991. Ueda, A., Ozono, R., Oshima, T., Yano, A., Kambe, M., Teranishi, Y., Katsuki, M., Chayama, K., 2003. Disruption of the type 2 dopamine receptor gene causes a sodium-dependent increase in blood pressure in mice. American Journal of Hypertension 16, 853–858. Usiello, A., Baik, J.H., Rouge-Pont, F., Picetti, R., Dierich, A., LeMeur, M., Piazza, P.V., Borrelli, E., 2000. Distinct functions of the two isoforms of dopamine D2 receptors. Nature 408, 199–203. Vallone, D., Pignatelli, M., Grammatikopoulos, G., Ruocco, L., Bozzi, Y., Westphal, H., Borrelli, E., Sadile, A.G., 2002. Activity, non-selective attention and emotionality in dopamine D2/D3 receptor knock-out mice. Behavioral Brain Research 130, 141–148. Van Tol, H.H., Bunzow, J.R., Guan, H.C., Sunahara, R.K., Seeman, P., Niznik, H.B., Civelli, O., 1991. Cloning of the gene for a human dopamine D4 receptor with high affinity for the antipsychotic clozapine. Nature 350, 610–614. Viggiano, D., Ruocco, L.A., Sadile, A.G., 2003. Dopamine phenotype and behaviour in animal models: in relation to attention deficit hyperactivity disorder. Neuroscience and Biobehavioral Reviews 27, 623–637. Volkow, N.D., Fowler, J.S., Wang, G.J., Goldstein, R.Z., 2002. Role of dopamine, the frontal cortex and memory circuits in drug addiction: insight from imaging studies. Neurobiology of Learning and Memory 78, 610–624. Vorel, S.R., Ashby Jr., C.R., Paul, M., Liu, X., Hayes, R., Hagan, J.J., Middlemiss, D.N., Stemp, G., Gardner, E.L., 2002. Dopamine D3 receptor antagonism inhibits cocaine-seeking and cocaine-enhanced brain reward in rats. The Journal of Neuroscience 22, 9595–9603. Vukhac, K.L., Sankoorikal, E.B., Wang, Y., 2001. Dopamine D2L receptor- and age-related reduction in offensive aggression. Neuroreport 12, 1035–1038.
1134
A. Holmes et al. / Neuropharmacology 47 (2004) 1117–1134
Waddington, J.L., Clifford, J.J., McNamara, F.N., Tomiyama, K., Koshikawa, N., Croke, D.T., 2001. The psychopharmacologymolecular biology interface: exploring the behavioural roles of dopamine receptor subtypes using targeted gene deletion (‘knockout’). Progress in Neuropsychopharmacology and Biological Psychiatry 25, 925–964. Wahlsten, D., Rustay, N.R., Metten, P., Crabbe, J.C., 2003. In search of a better mouse test. Trends in Neuroscience 26, 132–136. Walters, J.R., Bergstrom, D.A., Carlson, J.H., Chase, T.N., Braun, A.R., 1987. D1 dopamine receptor activation required for postsynaptic expression of D2 agonist effects. Science 236, 719–722. Wang, Y., Xu, R., Sasaoka, T., Tonegawa, S., Kung, M.P., Sankoorikal, E.B., 2000. Dopamine D2 long receptor-deficient mice display alterations in striatum-dependent functions. The Journal of Neuroscience 20, 8305–8314. Williams, G.V., Goldman-Rakic, P.S., 1995. Modulation of memory fields by dopamine D1 receptors in prefrontal cortex. Nature 376, 572–575. Wong, J.Y., Clifford, J.J., Massalas, J.S., Finkelstein, D.I., Horne, M.K., Waddington, J.L., Drago, J., 2003a. Neurochemical changes in dopamine D1, D3 and D1/D3 receptor knockout mice. European Journal of Pharmacology 472, 39–47. Wong, J.Y., Clifford, J.J., Massalas, J.S., Kinsella, A., Waddington, J.L., Drago, J., 2003b. Essential conservation of D1 mutant phenotype at the level of individual topographies of behaviour in mice lacking both D1 and D3 dopamine receptors. Psychopharmacology (Berlin) 167, 167–173. Xu, M., Hu, X.T., Cooper, D.C., Moratalla, R., Graybiel, A.M., White, F.J., Tonegawa, S., 1994a. Elimination of cocaineinduced hyperactivity and dopamine-mediated neurophysiological effects in dopamine D1 receptor mutant mice. Cell 79, 945–955. Xu, M., Moratalla, R., Gold, L.H., Hiroi, N., Koob, G.F., Graybiel, A.M., Tonegawa, S., 1994b. Dopamine D1 receptor mutant mice are deficient in striatal expression of dynorphin and in dopaminemediated behavioral responses. Cell 79, 729–742. Xu, M., Hu, X.T., Cooper, D.C., White, F.J., Tonegawa, S., 1996. A genetic approach to study mechanisms of cocaine action. Annals of the New York Academy of Sciences 801, 51–63. Xu, M., Koeltzow, T.E., Santiago, G.T., Moratalla, R., Cooper, D.C., Hu, X.T., White, N.M., Graybiel, A.M., White, F.J., Tonegawa, S., 1997. Dopamine D3 receptor mutant mice exhibit
increased behavioral sensitivity to concurrent stimulation of D1 and D2 receptors. Neuron 19, 837–848. Xu, M., Koeltzow, T.E., Cooper, D.C., Tonegawa, S., White, F.J., 1999. Dopamine D3 receptor mutant and wild-type mice exhibit identical responses to putative D3 receptor-selective agonists and antagonists. Synapse 31, 210–215. Xu, F., Gainetdinov, R.R., Wetsel, W.C., Jones, S.R., Bohn, L.M., Miller, G.W., Wang, Y.M., Caron, M.G., 2000a. Mice lacking the norepinephrine transporter are supersensitive to psychostimulants. Nature Neuroscience 3, 465–471. Xu, M., Guo, Y., Vorhees, C.V., Zhang, J., 2000b. Behavioral responses to cocaine and amphetamine administration in mice lacking the dopamine D1 receptor. Brain Research 852, 198–207. Xu, R., Hranilovic, D., Fetsko, L.A., Bucan, M., Wang, Y., 2002. Dopamine D2S and D2L receptors may differentially contribute to the actions of antipsychotic and psychotic agents in mice. Molecular Psychiatry 7, 1075–1082. Yarkov, A.V., Hanger, D., Reploge, M., Joyce, J.N., 2003. Behavioral effects of dopamine agonists and antagonists in MPTPlesioned D3 receptor knockout mice. Pharmacology, Biochemistry and Behavior 76, 551–562. Zahniser, N.R., Simosky, J.K., Mayfield, R.D., Negri, C.A., Hanania, T., Larson, G.A., Kelly, M.A., Grandy, D.K., Rubinstein, M., Low, M.J., Fredholm, B.B., 2000. Functional uncoupling of adenosine A(2A) receptors and reduced response to caffeine in mice lacking dopamine D2 receptors. The Journal of Neuroscience 20, 5949–5957. Zapata, A., Witkin, J.M., Shippenberg, T.S., 2001. Selective D3 receptor agonist effects of (+)-PD 128907 on dialysate dopamine at low doses. Neuropharmacology 41, 351–359. Zhang, J., Xu, M., 2001. Toward a molecular understanding of psychostimulant actions using genetically engineered dopamine receptor knockout mice as model systems. Journal of Addictive Diseases 20, 7–18. Zhang, D., Zhang, L., Lou, D.W., Nakabeppu, Y., Zhang, J., Xu, M., 2002a. The dopamine D1 receptor is a critical mediator for cocaine-induced gene expression. Journal of Neurochemistry 82, 1453–1464. Zhang, J., Zhang, D., Xu, M., 2002b. Identification of chronic cocaine-induced gene expression through dopamine d1 receptors by using cDNA microarrays. Annals of the New York Academy of Sciences 965, 1–9.