Skip to main content

Volume 3 Supplement 3

Special Issue: Towards Defining the Pathophysiology of Autistic Behavior

The neurobiology of mouse models syntenic to human chromosome 15q

Abstract

Autism is a neurodevelopmental disorder that manifests in childhood as social behavioral abnormalities, such as abnormal social interaction, impaired communication, and restricted interest or behavior. Of the known causes of autism, duplication of human chromosome 15q11–q13 is the most frequently associated cytogenetic abnormality. Chromosome 15q11–q13 is also known to include imprinting genes. In terms of neuroscience, it contains interesting genes such as Necdin, Ube3a, and a cluster of GABAA subunits as well as huge clusters of non-coding RNAs (small nucleolar RNAs, snoRNAs). Phenotypic analyses of mice genetically or chromosomally engineered for each gene or their clusters on a region of mouse chromosome seven syntenic to human 15q11–q13 indicate that this region may be involved in social behavior, serotonin metabolism, and weight control. Further studies using these models will provide important clues to the pathophysiology of autism. This review overviews phenotypes of mouse models of genes in 15q11–q13 and their relationships to autism.

Autism

Autism is a developmental brain disorder whose definition has been extended to autism spectrum disorders (ASDs), including non-pervasive development disorder and Asperger syndrome (Levitt and Campbell 2009). The prevalence of autism is increasing and currently stands at one in every 150 individuals. Its influence on modern society is increasing, especially in developed countries, where birthrates are low and the management of children with these disorders is a serious social issue. However, the cause of autism and its pathophysiology remain unknown and require further research from a number of different approaches. Among the neuropsychiatric diseases, autism is considered as an appropriate disorder that can be approached from biological aspects. This is partly because genetic studies show that the concordance rate of autism in identical twins is over 90%, which is higher than other neuropsychiatric diseases such as schizophrenia and mood disorders (Abrahams and Geschwind 2008).

Most autism cases are sporadic and their causes are still unknown; however, 10–20% have known causes, such as (1) comorbidity with other diseases, such as fragile X syndrome, tuberous sclerosis, and Rett syndrome; (2) genome abnormalities, including copy number variations (CNVs), such as 15q11–q13 and 16q11.2; and (3) single gene mutations such as NLGN3/4, NRXN1, and SHANK2/3 (Abrahams and Geschwind 2008; Berkel et al. 2010; Cook and Scherer 2008; Kumar et al. 2008; Marshall et al. 2008; Pinto et al. 2010; Toro et al. 2010; Vorstman et al. 2006; Weiss et al. 2008). Recent genetic screening using cutting-edge technologies has revealed more CNV cases than previously expected (Cook and Scherer 2008). Given the rapid progress in current technologies, such as array comparative genomic hybridization (array CGH) and next generation sequencers, and their availability for clinical use, it is likely that more CNV cases will be diagnosed in the near future.

CNVs are important not only in explaining human heterogeneity but also because they are responsible for several diseases, especially neuropsychiatric disorders including autism. Human chromosome 15q11–q13 is the second most frequent region to contain chromosome abnormalities, probably because it contains large repeat sequences. Depending on its copy variant, this region causes different diseases, such as Prader–Willi syndrome (PWS), Angelmann syndrome (AS), and autism, all sharing ASD features. Comparison of mouse models of 15q copy variants with phenotypes of genetically engineered mice of each gene, with or without imprinting, will provide some clues to the pathophysiology of autism.

Mouse chromosome seven is syntenic to human chromosome 15q11–q13

Chromosomal abnormalities were known as causes of autism before the era of current genome technology. Among them, human 15q11–q13 is the most notable chromosome region (Vorstman et al. 2006) (Fig. 1). This region is also known to be an imprinting region (Chamberlain and Lalande 2010a, b; Nicholls and Knepper 2001). Thus, there are three classified regions: paternally, maternally, and biallelically expressed regions. Human chromosome 15q11–q13 is conserved in mouse chromosome seven and it is believed that the imprinting status is conserved between mouse and human (Yang et al. 1998). Knockout (KO) mice for each gene corresponding to a human counterpart have been generated and studied (Table 1).

Fig. 1
figure 1

Genes on human chromosome 15q11–q13. Clusters of non-coding RNA, such as HBII-52 and HBII-85, are omitted. The original data are taken from UCSC Genome Browser (http://genome.ucsc.edu)

Table 1 Genes on human chromosome 15q11–q13 and the phenotypes of their genetically engineered mice

Paternally expressed genes

In the paternally expressed regions of the mouse chromosome, Mkrn3, Magel2, Necdin (Ndn), Snrpn-Snurf, and huge clusters of small nucleolar RNAs (snoRNAs) are present. Magel2 is involved in hypothalamic functions, and its deficient mice show abnormal circadian rhythm, body weight, and hormonal secretion, such as testosterone (Bischof et al. 2007; Kozlov et al. 2007; Mercer and Wevrick 2009). Ndn was first identified as a gene related to neural differentiation (Maruyama et al. 1991). Ndn KO mice were generated from multiple laboratories; however, they display different phenotypes, from no abnormality to respiratory distress and lethality, similar to PWS, depending on the strain of mouse (Gerard et al. 1999; Kuwako et al. 2005; Muscatelli et al. 2000; Tsai et al. 1999a). Muscatelli et al. observed a neonatal lethality of variable penetrance in Ndn mutants, and they showed a reduction in both oxytocin-producing and luteinizing hormone-releasing hormone (LHRH)-producing neurons in the hypothalamus and emotional and cognitive-related behaviors reminiscent of PWS (Muscatelli et al. 2000). Interestingly, oxytocin is known to be related to social behavior and is expected to be a candidate for therapeutics of ASD (Insel 2010). Ndn deficiency induces perinatal serotonergic alterations that affect the maturation and function of the respiratory network, inducing respiratory problems (Zanella et al. 2008). Serotonergic and catecholaminergic projections are reduced in the Ndn null embryo (Lee et al. 2005). Provided that serotonin is involved in symptoms of autism, Ndn is a candidate for further studies to reveal its involvement in abnormal social behavior. Additionally, from studies using Ndn KO mice, it is reported that Ndn is involved not only in TrkA signaling (Kuwako et al. 2005) but also promotes tangential migration of neocortical interneurons from the basal forebrain (Kuwajima et al. 2010). The fact that the neocortical interneurons are important for development of inhibitory neural circuits also makes Ndn a possible candidate for causing an imbalance of the excitatory/inhibitory ratio. Thus, although it is still controversial whether Ndn is a cause of PWS, it is an intriguing target gene for further analyses in terms of developmental brain disorders, including ASDs. Snrpn is a small nuclear ribonucleoprotein N and the Snrpn promoter is located in a maternally methylated CpG island within an imprinting center (IC). Mutations of Snrpn are not sufficient to induce PWS in mice, but a larger deletion including the PWS-IC and Ndn show phenotypes common to PWS infants (Yang et al. 1998). Paternal transmission of IC deletions results in a PWS-like phenotype in mice (Bielinska et al. 2000). The human PWS-IC could acquire a methylation imprint in mouse oocytes but the imprint is lost in somatic cells, suggesting that the human imprinting center demonstrates conserved acquisition, but diverged maintenance, of imprinting in a mouse model for AS imprinting defects (Johnstone et al. 2006). The main promoter of Snrpn is not required for the function of the PWS-IC, and the 4.8-kb deletion induces a severe phenotypic effect and a partial imprinting defect; however, this defect is less severe than that with the 42-kb deletion of the PWS-IC (Bressler et al. 2001; Yang et al. 1998). Paternal deletion from Snrpn to Ube3a in the mouse causes hypotonia, growth retardation, and partial lethality (Tsai et al. 1999b). The non-coding RNA region includes HBII-52 and HBII-85. Knockout mice of MBII-85, a mouse homolog of HBII-85, show growth retardation, increased anxiety, and hyperphagia (Ding et al. 2008; Skryabin et al. 2007). Recently, HBII-85 was reported as a cause of PWS (Duker et al. 2010; Sahoo et al. 2008). On the other hand, knockout mice of MBII-52, a mouse homolog of HBII-52, display increased RNA editing of the serotonin 2c receptor (5-HT2cR) and less locomotor activity (Doe et al. 2009). In terms of the relationship between serotonin and autism, MBII52 is an interesting candidate gene.

Maternally expressed genes

Maternally expressed genes in chromosome 15q11–q13 include Ube3a and Atp10a. Ube3a is an E3 ubiqutin ligase and is known to be a causative gene for AS (Kishino et al. 1997; Matsuura et al. 1997). Ube3a knockout mice show impaired motor function, inducible seizures, learning deficits (Jiang et al. 1998; Miura et al. 2002), and sleep disturbance (Colas et al. 2005). The ataxic and long potentiation phenotypes are rescued by reduction of αCaMKII inhibitory phosphorylation (van Woerden et al. 2007). Knock-in (KI) Ube3a-YFP fusion reporter mice show that Ube3a is imprinted in all regions of the brain, but is biallelically expressed in glial cells (Dindot et al. 2008). Maternally Ube3a-deficient mice show altered linking behavior, which has been associated with cerebellar control (Heck et al. 2008). Recently, Ube3a was extensively characterized in neurons and found to be involved in synaptic plasticity (Greer et al. 2010; Margolis et al. 2010; Sato and Stryker 2010; Yashiro et al. 2009). A genetic study also reported the involvement of Ube3a in ASD (Glessner et al. 2009). Ube3a is the primary candidate gene for causing the phenotypic features in duplication 15q autism. In contrast, Atp10a, an aminophospholipid translocase, has been less comprehensively studied than Ube3a, and its imprinting status remains unclear (DuBose et al. 2010; Kayashima et al. 2003). Mice with a radiation-induced maternal deletion, including Ube3a and Gabrb3, are a model of obesity and type two diabetes (Dhar et al. 2004). Mice with a large maternal deletion from Ube3a to Gabrb3, produced by chromosome engineering, show altered ultrasonic vocalization, impaired learning and memory, and display parent of origin effects (Jiang et al. 2010).

Non-imprinting genes

In the conventional biallelic region, there is a cluster of GABAA receptor subunits (Gabrb3, Gabra5, Gabrg3), Oca2, and Herc2. The GABAA receptor is one of the major neurotransmitter receptors and forms the most notable target in this region because an imbalance between excitatory and inhibitory neurons during development is a pathophysiological hypothesis of developmental brain disorders, including ASD (Rubenstein and Merzenich 2003). Of the GABAA receptor subunits, the GABAA β3 subunit (encoded by the Gabrb3 gene) is well studied. Its deficient mice show epileptic seizures, deficits in learning and memory, poor motor skills (DeLorey et al. 1998; Homanics et al. 1997), as well as sleep disturbance, such as lower REM sleep time and greater EEG delta power than the wild-type (WT) (Wisor et al. 2002). The level of norepinephrine in the spinal cord of Gabrb3 KO mice was less than that of WT mice, suggesting that β3 subunit-containing GABAA receptors mediate the trophic effects on a subpopulation of spinally projecting noradrenergic neurons (Ugarte et al. 2001). Furthermore, Gabrb3 N265M KI mice display altered general anesthetic actions (Jurd et al. 2003). Gabrb3 KO mice show expansion of the plexus of locus coeruleus dendrites, with a concurrent reduction in muscle tone and marble burying behavior (Hashemi et al. 2007). They also exhibit hypoplasia of cerebellar vermal lobules and impaired social and exploratory behaviors in addition to deficits in non-selective attention (DeLorey et al. 2008). Survivors of pan-neuronal Gabrb3 knockouts are overtly normal, whereas those of forebrain selective knockouts have reduced fertility, are hyperactive, and some become obese (Ferguson et al. 2007). An association between a specific GABAA β3 receptor defect and autism has been reported recently (Delahanty et al. 2010). Thus, Gabrb3 is also an intriguing candidate gene in the region with respect to autism. Gabra5 KO mice display enhanced spatial learning, decreased amplitude of inhibitory post-synaptic current (IPSC), and enhanced amplitude of paired-pulse facilitation of excitatory postsynaptic potential (EPSP) (Collinson et al. 2002). While Gabra5 KO mice exhibit a normal behavioral phenotype for fear-associated contextual learning (Cheng et al. 2006) and no differences in long-term potentiation (LTP) in hippocampal slices (Collinson et al. 2002), they show reduced threshold for the induction of LTP via mechanisms that are independent of synaptic inhibition (Martin et al. 2010). On the other hand, α5 (H105R) KI mice, in which a partial reduction of α5 GABAA receptor restricted to CA1-3 subfield of the hippocampus was achieved, exhibit facilitated trace fear conditioning (Crestani et al. 2002). Moreover, H105R mutant mice show attenuated prepulse inhibition (PPI) and increased spontaneous locomotor activity (Hauser et al. 2005). Herc2 contains a COOH-terminal HECT domain and functions as an assembly factor for the complex of an E3 ubiquitin ligase (Bekker-Jensen et al. 2010). Mutant mice of Herc2 show defects in growth (Lehman et al. 1998), but the physiological functions of Herc2 in the brain remain unknown.

Genes between BP1 and BP2

In addition to genes between breakpoints BP2 and BP3, there are four highly homologous genes (Nipa1, Nipa2, Cyfip1, and Tubgcp5) in between BP1 and BP2 (Chai et al. 2003), which appear to result in developmental behavioral problems when disturbed (Doornbos et al. 2009; Tam et al. 2010; van der Zwaag et al. 2010; von der Lippe et al. 2011). Cyfip1 (cytoplasmic FMRP interacting protein, also known as Sra-1) is an intriguing candidate among them because it binds with the Fragile X protein (Schenck et al. 2001). Microarray analysis using lymphoblastoid cells from males with autism due to a 15q11–q13 duplication reveal that Cyfip1 is upregulated (Nishimura et al. 2007). Although there is no mouse model of Cyfip1 at this moment, it would be interesting to make both Cyfip1 deletion and transgene animals since individuals with 15q duplications and AS deletion patients have a larger deletion that includes Cyfip1 (Sahoo et al. 2006).

Genomic diseases in human chromosome 15q11–q13

In humans, a paternal deletion of 15q11–q13 causes PWS in which the patient shows hypotonia, obesity, cognitive deficit, and behavioral problems, including obsessive-compulsive disorder (Cassidy and Driscoll 2009; Chamberlain and Lalande 2010b). Maternal deletion of this region causes AS, which includes motor dysfunction, frequent seizures, learning disability, absent speech, and characteristic happy demeanor (Van Buggenhout and Fryns 2009). Both PWS and AS share ASD features (Schanen 2006; Sutcliffe and Nurmi 2003). On the other hand, duplication of this region is known to be the most frequent cytogenetic abnormality in ASDs (Hogart et al. 2010; Vorstman et al. 2006). Autism with 15q duplication syndrome was first reported as partial trisomy of chromosome 15 (Gillberg et al. 1991). It is widely believed that maternally duplicated cases cause autism, but paternally duplicated ones do not (Bolton et al. 2001; Browne et al. 1997; Cook et al. 1997; Mao et al. 2000; Repetto et al. 1998; Roberts et al. 2002). On the other hand, cases with developmental delay are documented with paternal duplications (Bolton et al. 2004; Mao et al. 2000; Mohandas et al. 1999; Roberts et al. 2002; Veltman et al. 2005). A recent screening using multiplex ligation-dependent probe amplification (MLPA) did not detect maternal duplication, suggesting that ASD cases in maternal duplication are not as frequent as previously assumed (Depienne et al. 2009). Recent advances in genetic analyses, such as array CGH, and their clinical application are expected to detect larger numbers of clinical cases for 15q duplication.

A mouse model for 15q duplication

Using a chromosome-engineering technique (Mills and Bradley 2001; van der Weyden and Bradley 2006; Yu and Bradley 2001), a model for human chromosome 15q duplication has been generated (Nakatani et al. 2009). The duplication was confirmed in ES cells by Southern blotting and in mice by fluorescence in situ hybridization (FISH) and CGH arrays. Southern blot analysis using methylation sensitive and insensitive restriction enzymes, and a probe near the IC, indicated that allele-specific methylation is conserved in mice with the duplication. As the region is imprinted, the duplication is classified into two types: patDp/+ mice, in which the duplicated region is derived from the paternal allele; and matDp/+ mice, in which the duplication is derived from the maternal allele.

Both patDp/+ and matDp/+ mice breed normally and are fertile (Nakatani et al. 2009). The patDp/+ male mice begin to show an increase in body weight after 15 weeks and the body weight of patDp/+ mice is significantly greater than that of the WT after 20 weeks. The fact that both decreased (PWS) and increased (15q duplication) gene expressions of PWS regions lead to obesity is contradictory, but rather intriguing. As described above, Atp10a seems to be involved in metabolism and body weight (Dhar et al. 2004). Further, some of forebrain specific KO mice of Gabrb3 also become obese (Ferguson et al. 2007). The PWS region is involved in aspects of metabolism and above a certain critical threshold, the amounts of the genes and their products, regardless of any increase or decrease, may result in an abnormal pathway of metabolism.

Histological analyses, either at the macroscopic level or at the microscopic level, using simple staining such as hematoxylin–eosin or Bodian staining, detect no significant difference between mice with the duplication and WT mice (Nakatani et al. 2009). Recent studies reveal that synaptic adhesion molecules, and their associated proteins, are known to be rare mutation cases in autism (Betancur et al. 2009; Bourgeron 2009; Toro et al. 2010). For example, Shank3 mutant mice display abnormality in spines (Bozdagi et al. 2010; Peca et al. 2011). Thus, it will be quite interesting as to whether spine or synaptic abnormality is observed in 15q duplication mice or other models.

Autistic-like behavior of 15q duplication mice

Behavioral analyses are essential for animal models of autism because major phenotypes of ASDs are behavioral abnormalities and the diagnosis is made by three major phenotypes: impairment in social interaction, impaired social communication, restricted interest and behavior, or resistance to change. Mouse behavioral tests corresponding to human autistic phenotypes have been reported (Moy and Nadler 2008; Silverman et al. 2010). The systematic behavioral analyses were performed for 15q duplication mice (Nakatani et al. 2009). First, in the three-chamber test for social interaction, WT mice display more time spent near a novel mouse than an empty cage, while no difference of time spent around the novel mouse and the empty cage is observed in patDp/+ mice. In comparison with time between an inanimate object and an empty cage, both patDp/+ and WT mice spend more time near the object than empty cage, indicating that patDp/+ mice are able to see the object. Interestingly, WT mice spend more time around a novel mouse compared with an inanimate object, whereas there was no difference in time spent between the novel mouse and the inanimate object in patDp/+ mice. On the other hand, there was no significant difference between matDp/+ and WT mice in the three-chamber test. These results indicate that patDp/+ mice exhibit impaired social interaction. Similarly, Gabrb3−/− mice spend less time in the proximity of unfamiliar mice than do their control counterparts in a similar social behavior test, although some aspects of social curiosity appear to remain intact (DeLorey et al. 2008).

Second, patDp/+, but not matDp/+, mice display abnormal ultrasonic vocalizations (USVs) (Nakatani et al. 2009). When the pups are separated from their dams, they produce USVs. In WT pups, the numbers of USVs reach a peak around postnatal day five (P5) and then gradually decrease until P14, when their eyes are open and they can communicate through the visual pathway. In contrast, the numbers of USVs remain high at P7 or even at P14 in patDp/+ mice, suggesting that the developmental course of USVs in patDp/+ mice is delayed compared with WT. These results demonstrate that patDp/+ mice may be developmentally abnormal in comparison with WT, although the larger number of USVs in patDp/+ pups may reflect higher anxiety and fear in response to isolation stress. No difference of USVs between matDp/+ and WT mice was observed. The increased USVs were also observed in mice with a maternal deletion from Ube3a to Gabrb3 (Jiang et al. 2010).

Third, ritualistic behavioral inflexibility seen in autistic patients is examined by reversal tasks using learning mazes. In the Barnes maze, WT, patDp/+, and matDp/+ mice exhibit intact spatial learning and memory. In a reversal test, in which the target is altered to the opposite position, WT and matDp/+ mice learn the new target position, whereas patDp/+ mice return to the original position more often than WT and matDp/+ mice, indicating that patDp/+ mice display resistance to change or perseveration (Nakatani et al. 2009). patDp/+ mice also show similar behavior in the Morris water maze. These results suggest that patDp/+ mice exhibit behavioral inflexibility.

In addition to these phenotypes, patDp/+ mice display novelty-induced anxiety and decreased exploratory activity (Tamada et al. 2010). Interestingly, anxiety is also observed in paternal duplication of 15q11–q13 in humans (Mao et al. 2000; Veltman et al. 2005). Mice with maternal deficiency for just the Ube3a gene also show decreased exploratory behavior (Allensworth et al. 2011). In several tests involving food deprivation, such as the novelty suppressed test, radial maze, and Y-maze, patDp/+ mice display peculiar behavior (Tamada et al. 2010). Even though patDp/+ mice must be hungry after food deprivation, they do not rush to feed and take more time to proceed to feed than WT mice, and finally perform some tasks that are equivalent to the WT. This behavior is likely to reflect a lack of desire. patDp/+ mice also show less marble burying behavior (Tamada et al. 2010), a method commonly used to assess obsessive-compulsive behavior, which is also seen in Gabrb3 KO mice (Hashemi et al. 2007).

Abnormal serotonin in 15q duplication mice

The mouse model for 15q duplication can be considered an artificial founder mouse to unravel which gene and/or region in 15q11–q13 is critical for autistic phenotypes. A series of bacterial artificial chromosome (BAC), yeast artificial chromosome (YAC), or other engineered mice with longer insertions that cover the duplicated region will provide useful materials for further investigation. Alternatively, genetic rescue experiments crossed with a series of knockout or deficient mice will be required to identify the gene.

The primary benefit of the mouse model is that it allows us to study abnormality at the molecular and cellular levels. Major behavioral phenotypes appear in patDp/+ mice but not in matDp/+ mice (Nakatani et al. 2009); therefore, genes in the maternally imprinted region can be regarded as candidates for abnormal behaviors observed in patDp/+ mice. Among them, MBII-52, a mouse brain specific snoRNA 52 in a huge cluster of non-coding RNAs, would be an interesting candidate for the following reasons. First, MBII-52 is a paternally expressed gene. Second, HBII-52, a human homolog of MBII-52, is reported to be involved in posttranscriptional modification of the serotonin 2c receptor (5-HT2cR) (Kishore and Stamm 2006). Third, serotonin (5-HT) is well known to be related to the pathophysiology of autistic phenotypes (Whitaker-Azmitia 2005). Fourth, 5-HT2cR is the only G-protein coupled receptor that undergoes A to I RNA editing, resulting in an amino acid substitution in the second loop of the cytoplasmic domains (Seeburg 2002), which may affect G-protein coupling efficiency. MBII-52 RNA displays an approximately two-fold increase in expression in the patDp/+ mice brain compared with matDp/+ and WT mice (Nakatani et al. 2009). Pyrosequencing analyses revealed that the RNA editing ratio in potential editing sites of 5-HT2cR in patDp/+ mice was higher than that in matDp/+ or WT mice. This suggests that increased snoRNA in the patDp/+ brains may affect the RNA editing of 5-HT2cR. Furthermore, a specific agonist for 5-HT2cR induces higher intracellular Ca2+ concentration in primary cultured neurons derived from patDp/+ mice compared with matDp/+ and WT neurons, indicating that serotonergic signals in patDp/+ neurons are different from matDp/+ and WT neurons. This result suggests that alterations in RNA editing of 5-HT2cR may affect serotonergic signals in neurons and that this alteration in serotonergic signals in patDp/+ neurons may recapitulate abnormal behavior in patDp/+ mice.

Neurochemical analyses revealed abnormal levels of 5-HT and its metabolites in the adult cerebellum and midbrain of the duplication mouse. Interestingly, 5-HT in patDp/+ mice was also observed to decrease in all brain regions tested during postnatal weeks 1–3 (Tamada et al. 2010). These results suggest that abnormal 5-HT levels during developmental stages affect development of neurons and neural circuits, and may recapitulate the abnormal behavior seen in patDp/+ mice. Interestingly, the mean 5-HT concentrations are larger in Ndn KO mice than in WT, and the morphologies of 5-HT fibers are altered (Zanella et al. 2008). Combined with these findings, Ndn in the region may be involved in the metabolism and development of 5-HT. Ndn facilitates TrkA signaling for sensory neuron survival (Kuwako et al. 2005). The molecular mechanisms by which Ndn might affect 5-HT metabolism and development remain unknown.

Phenotypes (in)dependent on copy variants

Mice with a 15q duplication and other genetically engineered mice for genes in 15q11–q13 share some phenotypes, such as obesity, social behavior, and 5-HT concentration. Serotonin concentration seems to correlate with gene dosage of Ndn (Tamada et al. 2010; Zanella et al. 2008); however, complex phenotypes such as metabolism, weight control, or social behavior are unlikely to be dependent on simple gene dosage or copy number variants, but above a certain critical threshold, increases or decrease of gene expression may produce the abnormal pathway of these phenotypes.

The idea that phenotypes are dependent on copy variants may not be correct and their underlying mechanisms are unlikely to be simple in humans. In contrast to the interstitial duplication of human chromosome 15q, termed idic (15) syndrome individuals, the presence of a supernumerary marker chromosome formed by the inverted duplication of proximal chromosome 15, resulting in partial tetrasomy 15q, often display increased, yet inappropriate, social behaviors (Battaglia et al. 2010).

A gap between the mouse model and the human patient

The mouse model for 15q duplication as an animal model for human neuropsychiatric disease fulfills construct validity, in which it possesses the same chromosomal abnormality as human patients. Thus, the model can be regarded as a humanoid mouse model for autism in one aspect (Takumi 2010). In terms of humanized models, Tabuchi et al. report a KI mouse for an NLGN3 R541C mutant (Tabuchi et al. 2007). However, the same KI mouse independently generated by Dr. Heintz’s group displays no significant or mild social abnormality (Chadman et al. 2008). Further humanized model mice with rare mutations of other molecules, such as serotonin transporter G56A KI mice (Veenstra-Vanderweele et al. 2009), will be expected to display the phenotypes seen in humans.

The allele-specific phenotypes that appeared in 15q model mice will imply some biological significance that can provide a general concept. However, it is surprising that patDp/+ mice, but not matDp/+ mice, display abnormal social behavior (Nakatani et al. 2009) because maternally duplicated cases in humans are likely to result in an ASD phenotype, while paternal duplications do not (Cook et al. 1997). Although allele-specific methylation seems to be conserved, at least around the IC, in mice with a duplication, the possibility that the imprinting status in other regions is different between human and mouse is not excluded. The timing and mechanism of the expression of the downstream non-coding Snurf-Snrpn exons seem to be different between human and mouse (Chamberlain et al. 2010). Recent systematic analyses of the mouse epigenome reveal that Ube3a and a GABA receptor subunit are differentially imprinted in different brain areas (Gregg et al. 2010). Alternatively, epigenetic status other than DNA methylation may be different in mice from that in humans. Quantitative analysis of GABRB3 transcripts in PWS and AS brain samples show evidence for a paternal expression bias when a biparental 15q11–q13 contribution is lacking (Hogart et al. 2007). The Ube3a-ATS is thought to regulate expression by turning off Ube3a on the paternal allele. This may cause difference between maternal and paternal duplication animals as well and should be investigated. At the very least, there is a significant variability and diversity of clinical phenotypes and gene expression in human chromosome 15q11–q13 duplication syndrome (Hogart et al. 2009; Piard et al. 2010). Further detailed epigenetic analyses both in humans and mice will provide an insight into the imprinting mechanisms.

References

  • Abrahams BS, Geschwind DH. Advances in autism genetics: on the threshold of a new neurobiology. Nat Rev Genet. 2008;9:341–55.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Allensworth M, Saha A, Reiter LT, Heck DH. Normal social seeking behavior, hypoactivity and reduced exploratory range in a mouse model of Angelman syndrome. BMC Genet. 2011;12:7.

    PubMed Central  PubMed  Google Scholar 

  • Battaglia A, Parrini B, Tancredi R. The behavioral phenotype of the idic(15) syndrome. Am J Med Genet C Semin Med Genet. 2010;154C:448–55.

    PubMed  Google Scholar 

  • Bekker-Jensen S, Rendtlew Danielsen J, Fugger K, Gromova I, Nerstedt A, Lukas C, et al. HERC2 coordinates ubiquitin-dependent assembly of DNA repair factors on damaged chromosomes. Nat Cell Biol. 2010;12:80–6. sup pp 1–12.

    PubMed  CAS  Google Scholar 

  • Berkel S, Marshall CR, Weiss B, Howe J, Roeth R, Moog U, et al. Mutations in the SHANK2 synaptic scaffolding gene in autism spectrum disorder and mental retardation. Nat Genet. 2010;42:489–91.

    PubMed  CAS  Google Scholar 

  • Betancur C, Sakurai T, Buxbaum JD. The emerging role of synaptic cell-adhesion pathways in the pathogenesis of autism spectrum disorders. Trends Neurosci. 2009;32:402–12.

    PubMed  CAS  Google Scholar 

  • Bielinska B, Blaydes SM, Buiting K, Yang T, Krajewska-Walasek M, Horsthemke B, et al. De novo deletions of SNRPN exon 1 in early human and mouse embryos result in a paternal to maternal imprint switch. Nat Genet. 2000;25:74–8.

    PubMed  CAS  Google Scholar 

  • Bischof JM, Stewart CL, Wevrick R. Inactivation of the mouse Magel2 gene results in growth abnormalities similar to Prader–Willi syndrome. Hum Mol Genet. 2007;16:2713–9.

    PubMed  CAS  Google Scholar 

  • Bolton PF, Dennis NR, Browne CE, Thomas NS, Veltman MW, Thompson RJ, et al. The phenotypic manifestations of interstitial duplications of proximal 15q with special reference to the autistic spectrum disorders. Am J Med Genet. 2001;105:675–85.

    PubMed  CAS  Google Scholar 

  • Bolton PF, Veltman MW, Weisblatt E, Holmes JR, Thomas NS, Youings SA, et al. Chromosome 15q11–13 abnormalities and other medical conditions in individuals with autism spectrum disorders. Psychiatr Genet. 2004;14:131–7.

    PubMed  Google Scholar 

  • Bourgeron T. A synaptic trek to autism. Curr Opin Neurobiol. 2009;19:231–4.

    PubMed  CAS  Google Scholar 

  • Bozdagi O, Sakurai T, Papapetrou D, Wang X, Dickstein DL, Takahashi N, et al. Haploinsufficiency of the autism-associated Shank3 gene leads to deficits in synaptic function, social interaction, and social communication. Mol Autism. 2010;1:15.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Bressler J, Tsai TF, Wu MY, Tsai SF, Ramirez MA, Armstrong D, et al. The SNRPN promoter is not required for genomic imprinting of the Prader–Willi/Angelman domain in mice. Nat Genet. 2001;28:232–40.

    PubMed  CAS  Google Scholar 

  • Browne CE, Dennis NR, Maher E, Long FL, Nicholson JC, Sillibourne J, et al. Inherited interstitial duplications of proximal 15q: genotype–phenotype correlations. Am J Hum Genet. 1997;61:1342–52.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Cassidy SB, Driscoll DJ. Prader–Willi syndrome. Eur J Hum Genet. 2009;17:3–13.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Chadman KK, Gong S, Scattoni ML, Boltuck SE, Gandhy SU, Heintz N, et al. Minimal aberrant behavioral phenotypes of neuroligin-3 R451C knockin mice. Autism Res. 2008;1:147–58.

    PubMed Central  PubMed  Google Scholar 

  • Chai JH, Locke DP, Greally JM, Knoll JH, Ohta T, Dunai J, et al. Identification of four highly conserved genes between breakpoint hotspots BP1 and BP2 of the Prader–Willi/Angelman syndromes deletion region that have undergone evolutionary transposition mediated by flanking duplicons. Am J Hum Genet. 2003;73:898–925.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Chamberlain SJ, Lalande M. Angelman syndrome, a genomic imprinting disorder of the brain. J Neurosci. 2010a;30:9958–63.

    PubMed  CAS  Google Scholar 

  • Chamberlain SJ, Lalande M. Neurodevelopmental disorders involving genomic imprinting at human chromosome 15q11–q13. Neurobiol Dis. 2010b;39:13–20.

    PubMed  CAS  Google Scholar 

  • Chamberlain SJ, Chen PF, Ng KY, Bourgois-Rocha F, Lemtiri-Chlieh F, Levine ES, et al. Induced pluripotent stem cell models of the genomic imprinting disorders Angelman and Prader–Willi syndromes. Proc Natl Acad Sci USA. 2010;107:17668–73.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Cheng VY, Martin LJ, Elliott EM, Kim JH, Mount HT, Taverna FA, et al. Alpha5GABAA receptors mediate the amnestic but not sedative-hypnotic effects of the general anesthetic etomidate. J Neurosci. 2006;26:3713–20.

    PubMed  CAS  Google Scholar 

  • Colas D, Wagstaff J, Fort P, Salvert D, Sarda N. Sleep disturbances in Ube3a maternal-deficient mice modeling Angelman syndrome. Neurobiol Dis. 2005;20:471–8.

    PubMed  CAS  Google Scholar 

  • Collinson N, Kuenzi FM, Jarolimek W, Maubach KA, Cothliff R, Sur C, et al. Enhanced learning and memory and altered GABAergic synaptic transmission in mice lacking the alpha 5 subunit of the GABAA receptor. J Neurosci. 2002;22:5572–80.

    PubMed  CAS  Google Scholar 

  • Cook Jr EH, Scherer SW. Copy-number variations associated with neuropsychiatric conditions. Nature. 2008;455:919–23.

    PubMed  CAS  Google Scholar 

  • Cook Jr EH, Lindgren V, Leventhal BL, Courchesne R, Lincoln A, Shulman C, et al. Autism or atypical autism in maternally but not paternally derived proximal 15q duplication. Am J Hum Genet. 1997;60:928–34.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Crestani F, Keist R, Fritschy JM, Benke D, Vogt K, Prut L, et al. Trace fear conditioning involves hippocampal alpha5 GABA(A) receptors. Proc Natl Acad Sci USA. 2002;99:8980–5.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Delahanty RJ, Kang JQ, Brune CW, Kistner EO, Courchesne E, Cox NJ, et al. Maternal transmission of a rare GABRB3 signal peptide variant is associated with autism. Mol Psychiatry. 2010;16:86–96.

    Google Scholar 

  • DeLorey TM, Handforth A, Anagnostaras SG, Homanics GE, Minassian BA, Asatourian A, et al. Mice lacking the beta3 subunit of the GABAA receptor have the epilepsy phenotype and many of the behavioral characteristics of Angelman syndrome. J Neurosci. 1998;18:8505–14.

    PubMed  CAS  Google Scholar 

  • DeLorey TM, Sahbaie P, Hashemi E, Homanics GE, Clark JD. Gabrb3 gene deficient mice exhibit impaired social and exploratory behaviors, deficits in non-selective attention and hypoplasia of cerebellar vermal lobules: a potential model of autism spectrum disorder. Behav Brain Res. 2008;187:207–20.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Depienne C, Moreno-De-Luca D, Heron D, Bouteiller D, Gennetier A, Delorme R, et al. Screening for genomic rearrangements and methylation abnormalities of the 15q11–q13 region in autism spectrum disorders. Biol Psychiatry. 2009;66:349–59.

    PubMed  CAS  Google Scholar 

  • Dhar MS, Sommardahl CS, Kirkland T, Nelson S, Donnell R, Johnson DK, et al. Mice heterozygous for Atp10c, a putative amphipath, represent a novel model of obesity and type 2 diabetes. J Nutr. 2004;134:799–805.

    PubMed  CAS  Google Scholar 

  • Dindot SV, Antalffy BA, Bhattacharjee MB, Beaudet AL. The Angelman syndrome ubiquitin ligase localizes to the synapse and nucleus, and maternal deficiency results in abnormal dendritic spine morphology. Hum Mol Genet. 2008;17:111–8.

    PubMed  CAS  Google Scholar 

  • Ding F, Prints Y, Dhar MS, Johnson DK, Garnacho-Montero C, Nicholls RD, et al. Lack of Pwcr1/MBII-85 snoRNA is critical for neonatal lethality in Prader-Willi syndrome mouse models. Mamm Genome. 2005;16:424–31.

    PubMed  CAS  Google Scholar 

  • Ding F, Li HH, Zhang S, Solomon NM, Camper SA, Cohen P, et al. SnoRNA Snord116 (Pwcr1/MBII-85) deletion causes growth deficiency and hyperphagia in mice. PLoS One. 2008;3:e1709.

    PubMed Central  PubMed  Google Scholar 

  • Doe CM, Relkovic D, Garfield AS, Dalley JW, Theobald DE, Humby T, et al. Loss of the imprinted snoRNA mbii-52 leads to increased 5htr2c pre-RNA editing and altered 5HT2CR-mediated behaviour. Hum Mol Genet. 2009;18:2140–8.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Doornbos M, Sikkema-Raddatz B, Ruijvenkamp CA, Dijkhuizen T, Bijlsma EK, Gijsbers AC, et al. Nine patients with a microdeletion 15q11.2 between breakpoints 1 and 2 of the Prader–Willi critical region, possibly associated with behavioural disturbances. Eur J Med Genet. 2009;52:108–15.

    PubMed  Google Scholar 

  • DuBose AJ, Johnstone KA, Smith EY, Hallett RA, Resnick JL. Atp10a, a gene adjacent to the PWS/AS gene cluster, is not imprinted in mouse and is insensitive to the PWS-IC. Neurogenetics. 2010;11:145–51.

    PubMed  CAS  Google Scholar 

  • Duker AL, Ballif BC, Bawle EV, Person RE, Mahadevan S, Alliman S, et al. Paternally inherited microdeletion at 15q11.2 confirms a significant role for the SNORD116 C/D box snoRNA cluster in Prader–Willi syndrome. Eur J Hum Genet. 2010;18:1196–201.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Ferguson C, Hardy SL, Werner DF, Hileman SM, Delorey TM, Homanics GE. New insight into the role of the beta3 subunit of the GABAA-R in development, behavior, body weight regulation, and anesthesia revealed by conditional gene knockout. BMC Neurosci. 2007;8:85.

    PubMed Central  PubMed  Google Scholar 

  • Gardner JM, Nakatsu Y, Gondo Y, Lee S, Lyon MF, King RA, et al. The mouse pink-eyed dilution gene: association with human Prader-Willi and Angelman syndromes. Science. 1992;257:1121–4.

    PubMed  CAS  Google Scholar 

  • Gerard M, Hernandez L, Wevrick R, Stewart CL. Disruption of the mouse necdin gene results in early post-natal lethality. Nat Genet. 1999;23:199–202.

    PubMed  CAS  Google Scholar 

  • Gillberg C, Steffenburg S, Wahlstrom J, Gillberg IC, Sjostedt A, Martinsson T, et al. Autism associated with marker chromosome. J Am Acad Child Adolesc Psychiatry. 1991;30:489–94.

    PubMed  CAS  Google Scholar 

  • Glessner JT, Wang K, Cai G, Korvatska O, Kim CE, Wood S, et al. Autism genome-wide copy number variation reveals ubiquitin and neuronal genes. Nature. 2009;459:569–73.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Greer PL, Hanayama R, Bloodgood BL, Mardinly AR, Lipton DM, Flavell SW, et al. The Angelman Syndrome protein Ube3A regulates synapse development by ubiquitinating arc. Cell. 2010;140:704–16.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Gregg C, Zhang J, Weissbourd B, Luo S, Schroth GP, Haig D, et al. High-resolution analysis of parent-of-origin allelic expression in the mouse brain. Science. 2010;329:643–8.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Hashemi E, Sahbaie P, Davies MF, Clark JD, DeLorey TM. Gabrb3 gene deficient mice exhibit increased risk assessment behavior, hypotonia and expansion of the plexus of locus coeruleus dendrites. Brain Res. 2007;1129:191–9.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Hauser J, Rudolph U, Keist R, Mohler H, Feldon J, Yee BK. Hippocampal alpha5 subunit-containing GABAA receptors modulate the expression of prepulse inhibition. Mol Psychiatry. 2005;10:201–7.

    PubMed  CAS  Google Scholar 

  • Heck DH, Zhao Y, Roy S, LeDoux MS, Reiter LT. Analysis of cerebellar function in Ube3a-deficient mice reveals novel genotype-specific behaviors. Hum Mol Genet. 2008;17:2181–9.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Hogart A, Nagarajan RP, Patzel KA, Yasui DH, Lasalle JM. 15q11–13 GABAA receptor genes are normally biallelically expressed in brain yet are subject to epigenetic dysregulation in autism-spectrum disorders. Hum Mol Genet. 2007;16:691–703.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Hogart A, Leung KN, Wang NJ, Wu DJ, Driscoll J, Vallero RO, et al. Chromosome 15q11–13 duplication syndrome brain reveals epigenetic alterations in gene expression not predicted from copy number. J Med Genet. 2009;46:86–93.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Hogart A, Wu D, LaSalle JM, Schanen NC. The comorbidity of autism with the genomic disorders of chromosome 15q11.2–q13. Neurobiol Dis. 2010;38:181–91.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Homanics GE, DeLorey TM, Firestone LL, Quinlan JJ, Handforth A, Harrison NL, et al. Mice devoid of gamma-aminobutyrate type A receptor beta3 subunit have epilepsy, cleft palate, and hypersensitive behavior. Proc Natl Acad Sci USA. 1997;94:4143–8.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Insel TR. The challenge of translation in social neuroscience: a review of oxytocin, vasopressin, and affiliative behavior. Neuron. 2010;65:768–79.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Jiang YH, Armstrong D, Albrecht U, Atkins CM, Noebels JL, Eichele G, et al. Mutation of the Angelman ubiquitin ligase in mice causes increased cytoplasmic p53 and deficits of contextual learning and long-term potentiation. Neuron. 1998;21:799–811.

    PubMed  CAS  Google Scholar 

  • Jiang YH, Pan Y, Zhu L, Landa L, Yoo J, Spencer C, et al. Altered ultrasonic vocalization and impaired learning and memory in Angelman syndrome mouse model with a large maternal deletion from Ube3a to Gabrb3. PLoS One. 2010;5:e12278.

    PubMed Central  PubMed  Google Scholar 

  • Johnson DK, Stubbs LJ, Culiat CT, Montgomery CS, Russell LB, Rinchik EM. Molecular analysis of 36 mutations at the mouse pink-eyed dilution (p) locus. Genetics. 1995;141:1563–71.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Johnstone KA, DuBose AJ, Futtner CR, Elmore MD, Brannan CI, Resnick JL. A human imprinting centre demonstrates conserved acquisition but diverged maintenance of imprinting in a mouse model for Angelman syndrome imprinting defects. Hum Mol Genet. 2006;15:393–404.

    PubMed  CAS  Google Scholar 

  • Jong MT, Gray TA, Ji Y, Glenn CC, Saitoh S, Driscoll DJ, et al. A novel imprinted gene, encoding a RING zinc-finger protein, and overlapping antisense transcript in the Prader-Willi syndrome critical region. Hum Mol Genet. 1999;8:783–93.

    PubMed  CAS  Google Scholar 

  • Jurd R, Arras M, Lambert S, Drexler B, Siegwart R, Crestani F, et al. General anesthetic actions in vivo strongly attenuated by a point mutation in the GABA(A) receptor beta3 subunit. FASEB J. 2003;17:250–2.

    PubMed  CAS  Google Scholar 

  • Kayashima T, Ohta T, Niikawa N, Kishino T. On the conflicting reports of imprinting status of mouse ATP10a in the adult brain: strain-background-dependent imprinting? J Hum Genet. 2003;48:492–3. author reply 494.

    PubMed  Google Scholar 

  • Kishino T, Lalande M, Wagstaff J. UBE3A/E6-AP mutations cause Angelman syndrome. Nat Genet. 1997;15:70–3.

    PubMed  CAS  Google Scholar 

  • Kishore S, Stamm S. The snoRNA HBII-52 regulates alternative splicing of the serotonin receptor 2C. Science. 2006;311:230–2.

    PubMed  CAS  Google Scholar 

  • Kozlov SV, Bogenpohl JW, Howell MP, Wevrick R, Panda S, Hogenesch JB, et al. The imprinted gene Magel2 regulates normal circadian output. Nat Genet. 2007;39:1266–72.

    PubMed  CAS  Google Scholar 

  • Kumar RA, KaraMohamed S, Sudi J, Conrad DF, Brune C, Badner JA, et al. Recurrent 16p11.2 microdeletions in autism. Hum Mol Genet. 2008;17:628–38.

    PubMed  CAS  Google Scholar 

  • Kuwajima T, Hasegawa K, Yoshikawa K. Necdin promotes tangential migration of neocortical interneurons from basal forebrain. J Neurosci. 2010;30:3709–14.

    PubMed  CAS  Google Scholar 

  • Kuwako K, Hosokawa A, Nishimura I, Uetsuki T, Yamada M, Nada S, et al. Disruption of the paternal necdin gene diminishes TrkA signaling for sensory neuron survival. J Neurosci. 2005;25:7090–9.

    PubMed  CAS  Google Scholar 

  • Lee S, Walker CL, Karten B, Kuny SL, Tennese AA, O’Neill MA, et al. Essential role for the Prader–Willi syndrome protein necdin in axonal outgrowth. Hum Mol Genet. 2005;14:627–37.

    PubMed  CAS  Google Scholar 

  • Lehman AL, Nakatsu Y, Ching A, Bronson RT, Oakey RJ, Keiper-Hrynko N, et al. A very large protein with diverse functional motifs is deficient in rjs (runty, jerky, sterile) mice. Proc Natl Acad Sci USA. 1998;95:9436–41.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Levitt P, Campbell DB. The genetic and neurobiologic compass points toward common signaling dysfunctions in autism spectrum disorders. J Clin Invest. 2009;119:747–54.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Mao R, Jalal SM, Snow K, Michels VV, Szabo SM, Babovic-Vuksanovic D. Characteristics of two cases with dup(15)(q11.2–q12): one of maternal and one of paternal origin. Genet Med. 2000;2:131–5.

    PubMed  CAS  Google Scholar 

  • Margolis SS, Salogiannis J, Lipton DM, Mandel-Brehm C, Wills ZP, Mardinly AR, et al. EphB-mediated degradation of the RhoA GEF Ephexin5 relieves a developmental brake on excitatory synapse formation. Cell. 2010;143:442–55.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Marshall CR, Noor A, Vincent JB, Lionel AC, Feuk L, Skaug J, et al. Structural variation of chromosomes in autism spectrum disorder. Am J Hum Genet. 2008;82:477–88.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Martin LJ, Zurek AA, MacDonald JF, Roder JC, Jackson MF, Orser BA. Alpha5GABAA receptor activity sets the threshold for long-term potentiation and constrains hippocampus-dependent memory. J Neurosci. 2010;30:5269–82.

    PubMed  CAS  Google Scholar 

  • Maruyama K, Usami M, Aizawa T, Yoshikawa K. A novel brain-specific mRNA encoding nuclear protein (necdin) expressed in neurally differentiated embryonal carcinoma cells. Biochem Biophys Res Commun. 1991;178:291–6.

    PubMed  CAS  Google Scholar 

  • Matsuura T, Sutcliffe JS, Fang P, Galjaard RJ, Jiang YH, Benton CS, et al. De novo truncating mutations in E6-AP ubiquitin-protein ligase gene (UBE3A) in Angelman syndrome. Nat Genet. 1997;15:74–7.

    PubMed  CAS  Google Scholar 

  • Mercer RE, Wevrick R. Loss of magel2, a candidate gene for features of Prader–Willi syndrome, impairs reproductive function in mice. PLoS One. 2009;4:e4291.

    PubMed Central  PubMed  Google Scholar 

  • Mills AA, Bradley A. From mouse to man: generating megabase chromosome rearrangements. Trends Genet. 2001;17:331–9.

    PubMed  CAS  Google Scholar 

  • Miura K, Kishino T, Li E, Webber H, Dikkes P, Holmes GL, et al. Neurobehavioral and electroencephalographic abnormalities in Ube3a maternal-deficient mice. Neurobiol Dis. 2002;9:149–59.

    PubMed  CAS  Google Scholar 

  • Mohandas TK, Park JP, Spellman RA, Filiano JJ, Mamourian AC, Hawk AB, et al. Paternally derived de novo interstitial duplication of proximal 15q in a patient with developmental delay. Am J Med Genet. 1999;82:294–300.

    PubMed  CAS  Google Scholar 

  • Moy SS, Nadler JJ. Advances in behavioral genetics: mouse models of autism. Mol Psychiatry. 2008;13:4–26.

    PubMed  CAS  Google Scholar 

  • Muscatelli F, Abrous DN, Massacrier A, Boccaccio I, Le Moal M, Cau P, et al. Disruption of the mouse Necdin gene results in hypothalamic and behavioral alterations reminiscent of the human Prader–Willi syndrome. Hum Mol Genet. 2000;9:3101–10.

    PubMed  CAS  Google Scholar 

  • Nakatani J, Tamada K, Hatanaka F, Ise S, Ohta H, Inoue K, et al. Abnormal behavior in a chromosome-engineered mouse model for human 15q11–13 duplication seen in autism. Cell. 2009;137:1235–46.

    PubMed Central  PubMed  Google Scholar 

  • Nicholls RD, Knepper JL. Genome organization, function, and imprinting in Prader–Willi and Angelman syndromes. Annu Rev Genomics Hum Genet. 2001;2:153–75.

    PubMed  CAS  Google Scholar 

  • Nishimura Y, Martin CL, Vazquez-Lopez A, Spence SJ, Alvarez-Retuerto AI, Sigman M, et al. Genome-wide expression profiling of lymphoblastoid cell lines distinguishes different forms of autism and reveals shared pathways. Hum Mol Genet. 2007;16:1682–98.

    PubMed  CAS  Google Scholar 

  • Peca J, Feliciano C, Ting JT, Wang W, Wells MF, Venkatraman TN, et al. Shank3 mutant mice display autistic-like behaviours and striatal dysfunction. Nature. 2011;472:437–42.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Piard J, Philippe C, Marvier M, Beneteau C, Roth V, Valduga M, et al. Clinical and molecular characterization of a large family with an interstitial 15q11q13 duplication. Am J Med Genet A. 2010;152A:1933–41.

    PubMed  Google Scholar 

  • Pinto D, Pagnamenta AT, Klei L, Anney R, Merico D, Regan R, et al. Functional impact of global rare copy number variation in autism spectrum disorders. Nature. 2010;466:368–72.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Repetto GM, White LM, Bader PJ, Johnson D, Knoll JH. Interstitial duplications of chromosome region 15q11q13: clinical and molecular characterization. Am J Med Genet. 1998;79:82–9.

    PubMed  CAS  Google Scholar 

  • Roberts SE, Dennis NR, Browne CE, Willatt L, Woods G, Cross I, et al. Characterisation of interstitial duplications and triplications of chromosome 15q11–q13. Hum Genet. 2002;110:227–34.

    PubMed  CAS  Google Scholar 

  • Rubenstein JL, Merzenich MM. Model of autism: increased ratio of excitation/inhibition in key neural systems. Genes Brain Behav. 2003;2:255–67.

    PubMed  CAS  Google Scholar 

  • Sahoo T, Peters SU, Madduri NS, Glaze DG, German JR, Bird LM, et al. Microarray based comparative genomic hybridization testing in deletion bearing patients with Angelman syndrome: genotype–phenotype correlations. J Med Genet. 2006;43:512–6.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Sahoo T, del Gaudio D, German JR, Shinawi M, Peters SU, Person RE, et al. Prader–Willi phenotype caused by paternal deficiency for the HBII-85C/D box small nucleolar RNA cluster. Nat Genet. 2008;40:719–21.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Sato M, Stryker MP. Genomic imprinting of experience-dependent cortical plasticity by the ubiquitin ligase gene Ube3a. Proc Natl Acad Sci USA. 2010;107:5611–6.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Schaller F, Watrin F, Sturny R, Massacrier A, Szepetowski P, Muscatelli F. A single postnatal injection of oxytocin rescues the lethal feeding behaviour in mouse newborns deficient for the imprinted Magel2 gene. Hum Mol Genet. 2010;19:4895–905.

    PubMed  CAS  Google Scholar 

  • Schanen NC. Epigenetics of autism spectrum disorders. Hum Mol Genet. 2006;15(Spec No 2):R138–50.

    PubMed  CAS  Google Scholar 

  • Schenck A, Bardoni B, Moro A, Bagni C, Mandel JL. A highly conserved protein family interacting with the fragile X mental retardation protein (FMRP) and displaying selective interactions with FMRP-related proteins FXR1P and FXR2P. Proc Natl Acad Sci USA. 2001;98:8844–9.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Seeburg PH. A-to-I editing: new and old sites, functions and speculations. Neuron. 2002;35:17–20.

    PubMed  CAS  Google Scholar 

  • Silverman JL, Yang M, Lord C, Crawley JN. Behavioural phenotyping assays for mouse models of autism. Nat Rev Neurosci. 2010;11:490–502.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Skryabin BV, Gubar LV, Seeger B, Pfeiffer J, Handel S, Robeck T, et al. Deletion of the MBII-85 snoRNA gene cluster in mice results in postnatal growth retardation. PLoS Genet. 2007;3:e235.

    PubMed Central  PubMed  Google Scholar 

  • Sutcliffe JS, Nurmi EL. Genetics of childhood disorders: XLVII. Autism, part 6: duplication and inherited susceptibility of chromosome 15q11–q13 genes in autism. J Am Acad Child Adolesc Psychiatry. 2003;42:253–6.

    PubMed  Google Scholar 

  • Tabuchi K, Blundell J, Etherton MR, Hammer RE, Liu X, Powell CM, et al. A neuroligin-3 mutation implicated in autism increases inhibitory synaptic transmission in mice. Science. 2007;318:71–6.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Takumi T. A humanoid mouse model of autism. Brain Dev. 2010;32:753–8.

    PubMed  Google Scholar 

  • Tam GW, van de Lagemaat LN, Redon R, Strathdee KE, Croning MD, Malloy MP, et al. Confirmed rare copy number variants implicate novel genes in schizophrenia. Biochem Soc Trans. 2010;38:445–51.

    PubMed  CAS  Google Scholar 

  • Tamada K, Tomonaga S, Hatanaka F, Nakai N, Takao K, Miyakawa T, et al. Decreased exploratory activity in a mouse model of 15q duplication syndrome; implications for disturbance of serotonin signaling. PLoS One. 2010;5:e15126.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Toro R, Konyukh M, Delorme R, Leblond C, Chaste P, Fauchereau F, et al. Key role for gene dosage and synaptic homeostasis in autism spectrum disorders. Trends Genet. 2010;26:363–72.

    PubMed  CAS  Google Scholar 

  • Tsai TF, Armstrong D, Beaudet AL. Necdin-deficient mice do not show lethality or the obesity and infertility of Prader–Willi syndrome. Nat Genet. 1999a;22:15–6.

    PubMed  CAS  Google Scholar 

  • Tsai TF, Jiang YH, Bressler J, Armstrong D, Beaudet AL. Paternal deletion from Snrpn to Ube3a in the mouse causes hypotonia, growth retardation and partial lethality and provides evidence for a gene contributing to Prader–Willi syndrome. Hum Mol Genet. 1999b;8:1357–64.

    PubMed  CAS  Google Scholar 

  • Ugarte SD, Homanics GE, Hammond DL. Effect of embryonic knock-down of GABAA receptors on the levels of monoamines and their metabolites in the CNS of the mouse. Brain Res. 2001;904:290–7.

    PubMed  CAS  Google Scholar 

  • Van Buggenhout G, Fryns JP. Angelman syndrome (AS, MIM 105830). Eur J Hum Genet. 2009;17:1367–73.

    PubMed Central  PubMed  Google Scholar 

  • van der Weyden L, Bradley A. Mouse chromosome engineering for modeling human disease. Annu Rev Genomics Hum Genet. 2006;7:247–76.

    PubMed Central  PubMed  Google Scholar 

  • van der Zwaag B, Staal WG, Hochstenbach R, Poot M, Spierenburg HA, de Jonge MV, et al. A co-segregating microduplication of chromosome 15q11.2 pinpoints two risk genes for autism spectrum disorder. Am J Med Genet B Neuropsychiatr Genet. 2010;153B:960–6.

    PubMed Central  PubMed  Google Scholar 

  • van Woerden GM, Harris KD, Hojjati MR, Gustin RM, Qiu S, de Avila Freire R, et al. Rescue of neurological deficits in a mouse model for Angelman syndrome by reduction of alphaCaMKII inhibitory phosphorylation. Nat Neurosci. 2007;10:280–2.

    PubMed  Google Scholar 

  • Veenstra-Vanderweele J, Jessen TN, Thompson BJ, Carter M, Prasad HC, Steiner JA, et al. Modeling rare gene variation to gain insight into the oldest biomarker in autism: construction of the serotonin transporter Gly56Ala knock-in mouse. J Neurodev Disord. 2009;1:158–71.

    PubMed Central  PubMed  Google Scholar 

  • Veltman MW, Thompson RJ, Craig EE, Dennis NR, Roberts SE, Moore V, et al. A paternally inherited duplication in the Prader–Willi/Angelman syndrome critical region: a case and family study. J Autism Dev Disord. 2005;35:117–27.

    PubMed  Google Scholar 

  • von der Lippe C, Rustad C, Heimdal K, Rodningen OK. 15q11.2 microdeletion—seven new patients with delayed development and/or behavioural problems. Eur J Med Genet. 2011;54(3):357–60.

    PubMed  Google Scholar 

  • Vorstman JA, Staal WG, van Daalen E, van Engeland H, Hochstenbach PF, Franke L. Identification of novel autism candidate regions through analysis of reported cytogenetic abnormalities associated with autism. Mol Psychiatry. 2006;11(1):18–28.

    CAS  Google Scholar 

  • Walkowicz M, Ji Y, Ren X, Horsthemke B, Russell LB, Johnson D, et al. Molecular characterization of radiation- and chemically induced mutations associated with neuromuscular tremors, runting, juvenile lethality, and sperm defects in jdf2 mice. Mamm Genome. 1999;10:870–8.

    PubMed  CAS  Google Scholar 

  • Weiss LA, Shen Y, Korn JM, Arking DE, Miller DT, Fossdal R, et al. Association between microdeletion and microduplication at 16p11.2 and autism. N Engl J Med. 2008;358:667–75.

    PubMed  CAS  Google Scholar 

  • Whitaker-Azmitia PM. Behavioral and cellular consequences of increasing serotonergic activity during brain development: a role in autism? Int J Dev Neurosci. 2005;23:75–83.

    PubMed  CAS  Google Scholar 

  • Wisor JP, DeLorey TM, Homanics GE, Edgar DM. Sleep states and sleep electroencephalographic spectral power in mice lacking the beta 3 subunit of the GABA(A) receptor. Brain Res. 2002;955:221–8.

    PubMed  CAS  Google Scholar 

  • Yang T, Adamson TE, Resnick JL, Leff S, Wevrick R, Francke U, et al. A mouse model for Prader–Willi syndrome imprinting-centre mutations. Nat Genet. 1998;19:25–31.

    PubMed  CAS  Google Scholar 

  • Yashiro K, Riday TT, Condon KH, Roberts AC, Bernardo DR, Prakash R, et al. Ube3a is required for experience-dependent maturation of the neocortex. Nat Neurosci. 2009;12:777–83.

    PubMed Central  PubMed  CAS  Google Scholar 

  • Yu Y, Bradley A. Engineering chromosomal rearrangements in mice. Nat Rev Genet. 2001;2:780–90.

    PubMed  CAS  Google Scholar 

  • Zanella S, Watrin F, Mebarek S, Marly F, Roussel M, Gire C, et al. Necdin plays a role in the serotonergic modulation of the mouse respiratory network: implication for Prader–Willi syndrome. J Neurosci. 2008;28:1745–55.

    PubMed  CAS  Google Scholar 

Download references

Acknowledgments

The author acknowledges all members of the Takumi laboratory, especially Jin Nakatani for his persistent efforts in providing a useful mouse model, and thanks Kota Tamada for his preparation of the figure and table.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Toru Takumi.

Rights and permissions

Open Access This article is published under license to BioMed Central Ltd. This is an Open Access article is distributed under the terms of the Creative Commons Attribution License ( https://creativecommons.org/licenses/by/2.0 ), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Takumi, T. The neurobiology of mouse models syntenic to human chromosome 15q. J Neurodevelop Disord 3, 270–281 (2011). https://doi.org/10.1007/s11689-011-9088-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11689-011-9088-1

Keywords