Open access

Molecular Pathways of Down Syndrome Critical Region Genes

Written By

Ferdinando Di Cunto and Gaia Berto

Submitted: 01 May 2012 Published: 06 March 2013

DOI: 10.5772/53000

From the Edited Volume

Down Syndrome

Edited by Subrata Kumar Dey

Chapter metrics overview

2,619 Chapter Downloads

View Full Metrics

1. Introduction

1.1. Identification and annotation of the DSCR

Down syndrome (DS) is a very complex disorder that requires, even more than other human genetics diseases, a “system level” understanding [1,2], both under the clinical and under the molecular genetics perspectives. Under the clinical point of view, all individuals affected by Down syndrome are characterized by learning disabilities, distinctive facial features, and low muscle tone (hypotonia) in early infancy. However, in most cases the clinical picture is complicated by additional problems, such as heart defects, leukemia, and early-onset Alzheimer's disease [3,4]. The degree to which an individual is affected by these characteristics varies from mild to severe. After the pioneering description by J.L. Down in 1866, almost one century was needed to decipher the etiology of the syndrome. The work of Lejeune proved that DS was caused by an extra copy of chromosome 21 (HSA21) [5], thus providing the first evidence for a genetic basis of intellectual disability. The main implication of this seminal discovery is that the complex phenotype seen in DS patients [6] must be caused by overdosage of HSA21 genes. However, it also raised the outstanding questions of whether one or few HSA21 genes may play a dominant role in the syndrome and whether specific HSA21 genes could contribute to specific phenotypic tracts. Answering these questions is still of paramount importance, because the identification of one or few ‘dominant’ molecular players could pave the road for the development of targeted therapeutic approaches. The development of molecular karyotyping has provided strong support to the view that a restricted region of HSA21, commonly referred to as Down Syndrome Crtitical Region (DSCR) might be responsible for the different phenotypes that characterize DS. In 1976 Poissonnier and coworkers, by using chromosome staining methods, found that one DS patient not possessing an extra HSA21 had only a partial trisomy, involving 21q22.1 and 21q22.2 bands [7]. Afterwards, it turned out that partial trisomies are responsible for approximately 1% of DS cases [8,9]. These patients show variable phenotypes, depending on the extension of the triplicated region. Therefore, partial trisomies of genes carried by chromosome 21 have been extremely valuable in investigating the involvement in DS. The analysis of 10 partial trisomy patients, [10] suggested that two regions of chromosome 21 were linked to most of the Jackson signs [3], including cognitive disorders. These regions, referred to has DCR-1 and DCR-2, respectively, encompassed the 21q22.2 band and were located around the D21S55 Site Targeted Sequence (STS) and between D21S55 and the MX1 gene, respectively. Korenberg and coworkers studied a different population and observed that the proximal and distal regions of the 21q arm were also associated with the full DS phenotype [11]. Although these studies confirmed the strong association of DS phenotypes with the DCR-1 region, they also suggested that DS is a contiguous gene syndrome, arguing against a single DS chromosomal region responsible for most of the DS phenotypic features [11]. More recently, an additional causal link of the region located between D21S17 and ETS2 to clinical features of DS was confirmed through lattice analysis [12]. Although the notion of a DSCR has gained wide acceptance in DS research, it must be underscored that some of the data that support it remain controversial and that its existence has recently come under considerable question. Indeed, a detailed study of segmental trisomy 21 in DS subjects, performed by using array comparative genome hybridization (GCH), excludes the implication of a single but rather suggest that multiple regions of HSA21 contribute to many of the phenotypes of DS, including intellectual disability DSCR [13]. Despite these apparent inconsistencies, we think that, in practical terms, the crucial point is not to prove whether one or more “critical region” exist, but rather to understand which dosage-sensitive genes contribute to specific DS phenotypes. Indeed, it is quite clear that the classical “reductionist” approach of identifying one or few master genes, which has been very successful in the case of Mendelian disorders, is not appropriate to unravel the extremely more complicated case of DS. In this case, the overall phenotype is certainly produced by the combined action of several genes, causing complex rearrangements of different molecular networks [14]. The relevance of the mentioned studies has been to restrict the list of HSA21 genes that may contribute more significantly to the clinical manifestations.

For these motivations, in Tables 1 and 2 we adopt an inclusive definition of the DSCR, which extends from the RCAN1 gene to the MX1 gene. This definition takes into account not only the putative borders that have been identified in the mentioned studies, but also the fact that the RCAN1 gene as been commonly considered as part of the DSCR, even though a precise mapping on the current release of the human genome sequence (HG19) would locate it outside the centromeric border defined by [12]. Obviously, the usefulness of this information will strongly depend on the degree of functional characterization of the genes comprised in the interval. Under this respect, as it is generally true for the human genome, it must be recognized that our knowledge is still quite limited.

HSA21 was one of the first human chromosomes to be fully sequenced [15]. Nevertheless, the list of the possible functional sequences located in the DSCR has progressively changed, not only for the uncertainty of defining precise borders, but especially for the changes in the current view of what a human gene is. Obviously, the initial emphasis has been to identify the protein-coding sequences, whose number is approximately of 40, on the basis of a comprehensive definition of the DSCR and of the present annotation of the human genome (Table 1). However, systematic studies performed in the last few years revealed that many genomic sequences that have been initially considered as “junk DNA”, are endowed with extremely relevant functional potential [16]. Indeed, genome-wide interrogations have revealed that a large majority of the human genome is transcribed and that a significant proportion of transcripts appears to be non-protein coding (ncRNA). Although it is well recognized that some ncRNAs play essential enzymatic activities in translation, splicing and ribosome biogenesis, the functions of most ncRNAs are still unknown. It is now believed that they could participate in complex regulatory circuits responsible for the fine-tuning of gene expression at both the transcriptional and post-transcriptional levels [16]. The best known ncRNAs are miRNAs, ~22 nucleotide-long molecules that mediate post-transcriptional gene silencing by binding complementary sequences located in the 3’ UTR of the mRNAs. Long intergenic ncRNAs (lincRNA) represent a less characterized but more abundant and heterogeneous class, and comprise transcripts longer than 200 nt involved in many biological processes, including transcriptional control, epigenetic modification and post-transcriptional control on mRNAs [16]. A recent discovery demonstrated that both mRNAs and ncRNAs can deploy their functions by contributing to an extensive RNA-RNA interaction network, based on the competition of these molecules for the binding of shared miRNAs (the ceRNA hypothesis) [17-20]. Importantly, transcribed pseudogenes could also be involved in these complex regulatory interactions [21]. In light of this growing complexity, we think that the presence of many ‘non conventional’ sequences within the DSCR should be taken into consideration when exploring the molecular consequences of an increased dosage of this region. We provide an updated list of them in Table 2.

DCR Gene Name EntrezGene ID Main molecular function Essential references Expression in adult brain
1 RCAN1 1827 CaN inhibitor See main text Yes
1 CLIC6 54102 Channel See main text Yes
1 RUNX1 861 Transcription factor See main text Yes
1 SETD4 54093 Unknown No information Yes
1 CBR1 873 Enzyme [165] Yes
1 CBR3 874 Enzyme [165]
1 DOPEY2 9980 Unknown [166] Yes
1 MORC3 23515 RNA-binding [167]
1 CHAF1B 8208 Chromatin assembly [168] Yes
1 CLDN14 23562 Tight junctions component [169]
1 SIM2 6493 Transcription factor See main text Yes
1 HLCS 3141 Enzyme [170] Yes
1 DSCR6 53820 Unknown [171] Yes
1 PIGP 51227 Enzyme [172] Yes
1 TTC3 7267 E3 ligase See main text Yes
1 DSCR3 10311 Unknown [173] Yes
1 DYRK1A 1859 Protein kinase See main text Yes
1-2 KCNJ6 3763 Channel See main text
1-2 DSCR4 10281 Unknown [174]
1-2 DSCR8 84677 Unknown [175]
1-2 KCNJ15 3772 Channel [176]
1-2 ERG 2078 Transcription factor See main text Yes
1-2 ETS2 2114 Transcription factor See main text Yes
2 PSMG1 8624 Chaperone [177] Yes
2 BRWD1 54014 Transcription factor See main text Yes
2 HMGN1 3150 Transcription factor See main text Yes
2 WRB 7485 Protein trafficking [178] Yes
2 LCA5L 150082 Ciliary protein [179]
2 SH3BGR 6450 Unknown No information Yes
2 B3GALT5 10317 Enzyme [180]
2 C21orf88 114041 Unknown No information Yes
2 IGSF5 150084 Adhesion molecule [181]
2 PCP4 5121 Unknown [182] Yes
2 DSCAM 1826 Adhesion molecule [183]
2 BACE2 25825 Protease See main text Yes
2 FAM3B 54097 Cytokine [184]
2 MX2 4600 Unknown [185]
2 MX1 4599 Unknown [185] Yes

Table 1.

Summary of the protein-coding genes contained by the DSCR. The first column indicates whether the genes belong to the DCR-1, to the DCR-2 or to the overlap region. The evidence for expression in adult brain is derived from the EVOC data [186] contained in the Ensembl genome browser. Genes are given in their physical order, starting from the more centromeric sequence.

DCR Gene Name Ensembl ID EntrezGene ID HSA21 coordinates Gene Biotype Evidence of expression (EST)
1 LINC00160 ENSG00000230978 54064 36096105 - 36109478 lincRNA
1 AP000330.8 ENSG00000234380 100506385 36118054 - 36157183 Antisense
1 AF015262.2 ENSG00000234703 36508935 - 36511519 lincRNA +
1 RPL34P3 ENSG00000223671 54026 36844395 - 36844730 Pseudogene +
1 EZH2P1 ENSG00000231300 266693 36972030 - 36972320 Pseudogene
1 AF015720.3 ENSG00000230794 37085437 - 37105240 processed transcript +
1 MIR802 ENSG00000211590 768219 37093013 - 37093106 miRNA
1 RPS20P1 ENSG00000229761 54025 37097045 - 37097398 Pseudogene
1 PPP1R2P2 ENSG00000234008 54036 37259493 - 37260105 Pseudogene
1 AP000688.8 ENSG00000231106 37377636 - 37379899 lincRNA +
1 RPL23AP3 ENSG00000214914 8489 37388377 - 37388844 Pseudogene ++
1 RIMKLBP1 ENSG00000189089 54031 37422512 - 37423675 Pseudogene
1 AP000688.11 ENSG00000236677 37432730 - 37436706 Antisense +
1 U6 ENSG00000200213 1497008 37438843 - 37438950 snRNA
1 AP000688.14 ENSG00000230212 100133286 37441940 - 37498938 sense intronic
1 AP000688.15 ENSG00000236119 37455157 - 37462712 lincRNA +
1 AP000688.29 ENSG00000233393 37477179 - 37481988 lincRNA +
1 MEMO1P1 ENSG00000226054 728556 37502669 - 37504208 Pseudogene
1 CBR3-AS1 ENSG00000236830 100506428 37504065 - 37528605 lincRNA
1 RPS9P1 ENSG00000214889 8410 37504748 - 37505330 Pseudogene
1 RPL3P1 ENSG00000228149 8488 37541268 - 37542478 Pseudogene
1 Metazoa_SRP ENSG00000265882 37585858 - 37586136 miscellaneous RNA
1 snoU13 ENSG00000238851 37630724 - 37630829 snoRNA
1 SRSF9P1 ENSG00000214867 54021 37667471 - 37668000 Pseudogene
1 AP000692.9 ENSG00000228107 37732928 - 37734338 processed transcript +
1 ATP5J2LP ENSG00000224421 54100 37761176 - 37761410 Pseudogene
1 AP000695.6 ENSG00000230479 37802658 - 37853368 Antisense +
1 AP000695.4 ENSG00000233818 37818029 - 37904706 Antisense
1 PSMD4P1 ENSG00000223741 54035 37858281 - 37859709 Pseudogene +
1 AP000696.2 ENSG00000231324 38004979 - 38009331 lincRNA ++
1 AP000697.6 ENSG00000224269 38071073 - 38073864 Antisense +
1 HLCS-IT1 ENSG00000237646 100874294 38176285 - 38178585 sense intronic ++
1 RN5S491 ENSG00000199806 100873733 38224211 - 38224328 rRNA
1 AP000704.5 ENSG00000224790 38338812 - 38344128 lincRNA ++
1 Y_RNA ENSG00000207416 38359039 - 38359151 miscellaneous RNA
1 MRPL20P1 ENSG00000215734 359737 38366943 - 38367375 Pseudogene
1 U6 ENSG00000212136 1497008 38417830 - 38417936 snRNA
1 TTC3-AS1 ENSG00000228677 100874006 38559967 - 38566227 Antisense ++
1 DSCR9 ENSG00000230366 257203 38580804 - 38594037 lincRNA
1 Metazoa_SRP ENSG00000263969 38587906 - 38588202 miscellaneous RNA
1 AP001432.14 ENSG00000242553 38593720 - 38610045 lincRNA +
1-2 KCNJ6-IT1 ENSG00000233213 100874329 39089405 - 39091872 sense intronic +
1-2 AP001427.1 ENSG00000264691 39334968 - 39335068 miRNA +
1-2 DSCR4-IT1 ENSG00000223608 100874327 39378846 - 39382920 sense intronic +
1-2 snoU13 ENSG00000238581 39559551 - 39559656 snoRNA
1-2 DSCR10 ENSG00000233316 259234 39578250 - 39580738 lincRNA
1-2 AP001434.2 ENSG00000226012 39609139 - 39610123 lincRNA +
1-2 SPATA20P1 ENSG00000231123 100874060 39610149 - 39610586 Pseudogene
1-2 AP001422.3 ENSG00000231231 39695557 - 39705343 lincRNA ++
1-2 SNRPGP13 ENSG00000231480 100874428 39874369 - 39874545 Pseudogene
1-2 LINC00114 ENSG00000223806 400866 40110825 - 40140898 lincRNA
2 AP001042.1 ENSG00000229986 40218171 - 40220568 lincRNA
2 AF064858.6 ENSG00000205622 400867 40249215 - 40328392 lincRNA
2 AP001043.1 ENSG00000229925 40260696 - 40275829 processed transcript +
2 SNORA62 ENSG00000252384 40266709 - 40266791 snoRNA
2 RPSAP64 ENSG00000227721 40266841 - 40267176 Pseudogene
2 AP001044.2 ENSG00000234035 40285093 - 40287072 lincRNA +
2 AF064858.7 ENSG00000232837 40346355 - 40349700 lincRNA +
2 AF064858.8 ENSG00000235888 40360633 - 40378079 lincRNA +
2 AF064858.11 ENSG00000237721 40378574 - 40383255 lincRNA +
2 AF064858.10 ENSG00000237609 40400461 - 40401053 lincRNA +
2 RPL23AP12 ENSG00000228861 391282 40499494 - 40499966 Pseudogene +
2 PCBP2P1 ENSG00000235701 54040 40543056 - 40544032 Pseudogene
2 TIMM9P2 ENSG00000232608 100862727 40588550 - 40589432 Pseudogene
2 BRWD1-IT1 ENSG00000237373 40589019 - 40591731 processed transcript +
2 METTL21AP1 ENSG00000229623 100421629 40607312 - 40607946 Pseudogene
2 BRWD1-AS1 ENSG00000238141 100874093 40687633 - 40695144 Antisense +
2 Y_RNA ENSG00000252915 40716463 - 40716554 miscellaneous RNA
2 snoU13 ENSG00000238556 40717300 - 40717383 snoRNA
2 RNF6P1 ENSG00000227406 100420924 40745689 - 40748992 Pseudogene
2 MYL6P2 ENSG00000235808 100431168 40860253 - 40860686 Pseudogene ++
2 RPS26P4 ENSG00000228349 692146 40863470 - 40863824 Pseudogene +
2 AF121897.4 ENSG00000235012 40897510 - 40901782 Pseudogene
2 AF064860.5 ENSG00000225330 41002198 - 41098012 processed transcript +
2 AF064860.7 ENSG00000231713 41099682 - 41102607 lincRNA +
2 MIR4760 ENSG00000263973 100616148 41584279 - 41584358 miRNA
2 DSCAM-AS1 ENSG00000235123 100506492 41755010 - 41757285 Antisense
2 SNORA51 ENSG00000207147 41885071 - 41885206 snoRNA
2 AF064863.1 ENSG00000221396 41949429 - 41949538 miRNA +
2 DSCAM-IT1 ENSG00000233756 100874326 41987304 - 42002693 sense intronic ++
2 YRDCP3 ENSG00000230859 100861429 42235920 - 42236399 Pseudogene
2 LINC00323 ENSG00000226496 284835 42513427 - 42520060 Antisense
2 MIR3197 ENSG00000263681 100423023 42539484 - 42539556 miRNA
2 AL773572.7 ENSG00000225745 42548249 - 42558715 processed transcript ++
2 BACE2-IT1 ENSG00000224388 282569 42552024 - 42552553 Antisense +
2 AP001610.5 ENSG00000228318 42813321 - 42814669 Antisense +

Table 2.

Summary of the non-protein-coding elements contained by the DSCR. The first column indicates whether the genes belong to the DCR-1, to the DCR-2 or to the overlap region. Elements are given in their physical order, starting from the more centromeric sequence. Genomic coordinates refer to the HG19 version of the human genome sequence. The evidence for expression is derived from the ESTs linked to the Ensembl genome browser. + = at least one EST sequence supporting the Ensemble prediction. ++ prediction supported by several EST sequences.

Advertisement

2. Functional analysis of the DSCR through mouse models

Animal models are essential to understand the molecular pathogenesis of DS. Moreover, although none of them can faithfully mimic the human situation, they are crucial for the preclinical development of new therapeutic strategies. The availability of sophisticated tools for mouse genetics and the conserved synteny between mouse chromosome 16 (MMU16) and HSA21 have provided the basis for the development of many mouse models of DS, allowing to test the critical region concept and to perform a genetic dissection of the complex DS phenotype.

The first mouse models have been obtained by studying the effects of partial trisomies of MMU16 derived from Robertsonian translocations. These mice live until adulthood and show many clinical phenotypes similar to DS patients, in particular the neuropathological and neurobiological alterations, including learning and behavioral abnormalities [22-25]. The most studied mouse model for DS is theTs65Dn mouse, which possesses an extra copy of the distal 13 Mbp part of MMU16, including ~ 104 mouse genes orthologous to those on HSA21 [23]. These mice show a number of developmental and functional parallels with DS, including craniofacial abnormalities and behavioural changes [26-32]. Moreover, they show alterations in the structure of dendritic spines in cortex and hippocampus [33] and reduced long-term potentiation (LTP) in the hippocampus and fascia dentata (FD) [34-36].

Ts1Cje mice, which are trisomic for a shorter but fully overlapping segment of MMU16 (~81 genes), show similar changes, usually to a lesser degree [24,25,37,38]. Comparison of the behavioral performances of the Ts1Cje and Ts65Dn showed that the learning deficits of Ts1Cje mice are similar to those of Ts65Dn. The data obtained from these models strongly supported the concept of DSCR, because they indicated that conserved genes are capable to influence cognition through their dosage lie in a region spanning from Sod1 to Mx1, which contains the mouse counterpart of the human DCR-1.

Probably, the most elegant studies that have addressed the role of the mouse genome region syntenic to the human DSCR are those undertaken by Roger H. Reeves and coworkers. Using chromosome engineering, this group has generated a mouse line referred to as Ts1Rhr, trisomic for a segment closely corresponding to the DCR-1 region, as defined by [10] and [11] and including 33 genes [39]. Moreover, they obtained the corresponding deletion, resulting in the monosomic line Ms1Rhr. Interestingly, the first results produced by the analysis of these models did not confirm strongly the DSCR hypothesis. Indeed, the craniofacial dysmorphologies of Ts1Rhr are less marked and distinct from those detected in Ts65Dn and Ts1Cje mice [39]. Furthermore, no differences were initially detected between Ts1Rhr and normal controls in the Morris water maze, in the induction of LTP in the hippocampal CA1 Region and in the hippocampal and in cerebellum volume [39-41]. These results seemed to suggest that triplication of the Ts1Rhr segment is not sufficient to produce these correlates of DS phenotypes. However, the intercross of the monosomic line Ms1Rhr with the Ds65Dn line, which restored in a disomic condition for DCR-1 genes, generated mice showing normal performances in the Morris water maze, indicating that trisomy of DCR-1 is necessary for these cognitive phenotypes [41]. Importantly, a more recent report established that, if the Ts1Rhr mutation is analyzed on the same genetic background of the Ts65Dn and Ts1Cje mice and with more stringent tests, important cognitive and synaptic neurobiological phenotypes can be detected [42]. In particular, 20 of 48 phenotypes, many of which are shared with Ts65Dn mice, distinguished Ts1Rhr animals from their 2N controls. In addition to the genetic background difference, it must be noticed that the task used in this work was less stressful and more sensitive than the water maze, which may further account for the initial discrepancy [42]. These phenotypes were correlated with changes in synaptic density and in dendritic spine morphology, further indicating that DCR-1 genes strongly contribute to these abnormalities [42]. In conclusion, taken together, these results provide strong support to the view that increased dosage of DCR1 genes is necessary and sufficient to confer to mice some of the neurobiological phenotypes characteristic of DS.

The use of mouse genetic tools has allowed the production of even more restricted models, addressing the role of specific subregions of the human or mouse DSCR, or even the role of single DSCR genes. For instance, the isolation from the DSCR of huge genomic clones maintained as Yeast Artificial Chromosomes (YAC) or as Bacterial Artificial Chromosomes (BAC) and their microinjection in mouse oocytes has allowed the generation of transgenic lines covering the entire length of the human DSCR [43-45]. The characterization of these mice has shown that the approach can be very useful to study the function of specific genes. However, it became also clear that this strategy is of limited usefulness to establish genes contribution to the phenotype. For instance, BAC transgenesis allowed the production of a mouse line carrying a single extra copy of the DYRK1A gene [46]. Interestingly, these mice showed impaired cognitive behaviours, but they were characterized by increased hippocampal LTP, while all the models discussed above show depressed hippocampal LTP [46]. The same conclusion applies even better to the models obtained through classical transgenesis approaches, in which a single human or mouse gene is inserted in the mouse genome in the form of a cDNA driven by a non-physiological promoter [47].

On the other hand, the combination of gene targeting technologies with the “classical” DS model discussed above allows a subtractive strategy, providing the most stringent test to address the relevance of single genes for the overall phenotype. Indeed, once a null allele for a DSCR gene is available, a compound mutant can be generated, carrying the specific mutation in a trisomic background. The subtractive approach allowed to detect a significant rescue of the phenotype in the case of some DS-related genes, belonging to the DSCR as in the case of DSCR1[48], Olig1 and Olig2 [49], or even external to it, as in the case of APP [50,51].

Advertisement

3. Functional role of DSCR genes in DS intellectual disability: Towards the identification of drugable pathways

In the following section we will summarize the most relevant functional information available on DSCR genes, trying to especially underscore their implication in molecular networks relevant to intellectual disability. As it is obvious from the previous sections, this discussion will involve not only genes that strictly belong to the DSCR, but also their interactions with other HSC21 genes, whose functional involvement is supported by abundant literature. In particular, we will try to discuss as much as possible the single DSCR genes on the basis of their common features. The essential information about genes not included in this section is reported in Tables 1 and 2. While deploying this summary, we will also provide a perspective of how this information can be useful for progressing towards the development of new therapeutic strategies that may take into account the complex nature of DS.

3.1. Pathogenesis of intellectual disability in DS

In order to evaluate the possible degree of functional involvement for specific genes, it is very important to briefly analyze the principal biological processes that have been to cognitive impairment in the DS. To this regard, studies performed both in humans and in animal models have shown that trisomy 21 leads to an unbalance of key cellular events, such as neuronal cell proliferation and differentiation, which can be detected during development and post-natal life using morphological methods [52,53]. Importantly, these defects may coexist with or may be causally related to functional deficits, that can be revealed using sophisticated physiological methods [52,53]. Reduced neurons number is found in cortex, hippocampus and cerebellum of DS brain and are accompanied by impaired neuronal function. Brain hypocellularity is acquired during early developmental stages and is paralleled by impaired cognitive development leading to intellectual disabilities. Further deterioration of cognitive abilities occurs in adolescence and adulthood, possibly due to degenerative mechanisms [28]. Although the syndrome invariably results in AD-like neuropathology, the actual onset of dementia is quite variable. The availability of genetic models of trisomy 21 has been instrumental in gaining insights into the pathogenic mechanisms leading to DS cognitive disability. Morphological abnormalities of neuronal dendritic compartment are paralleled by functional electrophysiological deficits and impairment of learning and memory, pointing to the existence of defective neural network connectivity and faulty neuronal communication as primary determinants of DS cognitive disabilities [34-38,42,54]. Such pathological scenario arises from a combination of neurodevelopmental abnormalities and neurodegenerative processes. Addressing which processes are irreversible and which ones can be prevented or reverted by manipulating genes and pathways is of paramount importance for the development of new therapeutic strategies. Although the crossover between neurogenesis dysfunction and neurodegeneration is still poorly understood, it is likely that common pathways differentially affect various cellular functions during development and aging. Thus, the developmental aspects are fundamental in defining the most important functional consequences of the genetic imbalance in DS at the cognitive level. However, the IQ of DS patients decreases in the first decade of life, indicating that the maturation of central nervous system is compromised [8]. Indeed, on one side, different observations suggest that neurogenesis impairment starting from the earliest stages of development may underlie the widespread brain atrophy of DS, the delayed and disorganized lamination in the DS fetal cortex [55] and hippocampal hypoplasia [56]. On the other, postmortem studies show that DS patients start their lives with an apparently normal neuronal architecture that progressively degenerates. During the peak period of dendritic growth and differentiation (2.5 months old infants), no significant differences were detected in dendritic differentiation between euploid and DS cases in pyramidal neurons of prefrontal cortex [57]. Similarly, DS infants younger than 6 months showed greater dendritic branching and length than normal infants [58] [59] in contrast to the reduced number of dendrites and degenerative changes in DS children older than two years [60].

3.2. Transcription factors and co-factors encoded by the DSCR

The DSCR contains 6 genes encoding for transcription factors (Table 1), which are likely to play crucial roles in determining DS phenotypes, considering their potential to affect many cellular networks. Two of them, ERG and ETS2 belong to the erythroblast transformation-specific (ETS) family. Members of this family are key regulators of embryonic development, cell proliferation, differentiation, angiogenesis, inflammation, and apoptosis [61]. ERG is required for vascular cell remodeling and hematopoesis [62,63], while ETS2 has been linked to thymocytes development and apoptosis [64]. Together with RUNX1 [65], these proteins are very likely to contribute to the hematological abnormalities that characterize DS, but not to contribute significantly to ID. In contrast, BRWD1 and HGMN1 are two proteins highly expressed in brain that is involved in chromatin-remodeling [66,67]. Importantly, HGMN1 has been found to regulate the expression of the ID gene MeCP2 [67]. Under the same perspective, another interesting candidate is the bHLH factor SIM2 that together with its paralog SIM1 is the homolog of Drosophila single-minded (sim) gene. The Drosophila sim gene encodes a transcription factor that is a master regulator of fruit fly neurogenesis [68], raising the possibility that SIM2 could perform a similar function in mammals. However, a role of SIM2 in mammalian neurogenesis has not been so far confirmed, while this gene has been shown to repress myogenesis in mouse [69]. Besides to directly regulating transcription, DSCR genes could strongly modulate the activity of transcription factors encoded by other loci. The best characterized example is RCAN1, which was initially named DSCR1 [70]. The gene name was then changed after realizing that the encoded protein inhibits calcineurin-dependent transcriptional responses by binding to the catalytic domain of calcineurin A and interfering with the phosphorylation of the NFAT transcription factor [71,72]. RCAN1 is overexpressed in DS brain [14,73] and seems to play a key role in the regulation of mitochondrial function and oxidative stress. Indeed, the Drosophila homolog of RCAN1 especially affects the activity of the mitochondrial ADP/ATP translocator [74]. Moreover, it has been shown that, when RCAN1 is overexpressed in PC12 cells, it induces the expression of superoxide dismutase type 1 (SOD1) [75], which is encoded by another HSA21 gene [15] and is upregulated in DS brain [76]. Importantly, RCAN1 acts as a stress response element: its acute overexpression protects cells from oxidative stress [77]. Indeed, RCAN1 overexpression may have beneficial effects by counteracting the oxidative damage associated with DS. Elevated levels of DNA damage, lipid peroxidation [78] and pro-oxidant state develop early in life in DS subjects [79]. Nevertheless, it is very likely that the benefits arising from these actions on oxidative stress may be overcome by the long-term detrimental effects on synaptic functions and neuronal survival due to the chronic RCAN1 overexpression, which will be discussed in sections 3.4 and 3.5.

3.3. Signaling proteins encoded by the DSCR

Modifications of the cellular cytoskeleton in response to extracellular stimuli, such as growth factor engagement and cell-cell contacts are essential for neuronal proliferation, for the formation of axons and Dendrites, for the differentiation and for the establishment, maintenance and remodeling of neuronal connections. Many of the well-characterized DSCR genes, such as DSCAM, CLDN14, PIGP, LCA5L, IGSF5 and FAM3B are implicated in these processes. However, the best characterized proteins belonging to this category are DYRK1A and TTC3.

3.3.1. DYRK1A

DYRK1A, dual-specificity tyrosine-phosphorilation-regulated kinase1A, encodes a protein kinase capable to phosphorylate serine, threonine and tyrosine residues, highly conserved at the aminoacidic level across vertebrates and invertebrates [80]. The orthologus Drosophila gene is involved in neuroblast proliferation and it is named minibrain (MNB), because null mutations affect post-embrionic neurogenesis, resulting in reduced brain size [81]. The highly conserved structure of this kinase and its mapping to the DSCR prompted extensive studies on its vertebrate homologues [82]. These studies have revealed that the dosage of DYRK1A is extremely important to normal brain development. Indeed, mice homozygous for a null mutation of DYRK1A die early in development and even heterozygous mice display reduced viability and a smaller brain, characterized by reduction of neuronal counts in specific regions [83]. Accordingly, truncation of the human MNB⁄DYRK1A gene has been reported to cause microcephaly [84,27]. Furthermore transgenic mice overexpressing DYRK1A show severe impairment in spatial learning and memory in the Morris water maze tests, indicating hippocampal and prefrontal cortical function alteration [45,85]. Moreover, these transgenic mice show abnormal LTP and LTD, indicating synaptic plasticity alterations [46]. These defects are similar to those found in murine models of DS with trisomy of chromosome 16, suggesting a causative role of DYRK1A in cognitive disorders present in DS patients. DYRK1A is expressed in the cortex, in the hippocampus and in the cerebellum [86,18] and is overexpressed in the mouse trisomic model Ts65Dn [87], in DS fetal brain and other trisomic tissues [88]. These data obtained from different experimental systems have revealed various possible functions of DYRK1A in central nervous system (CNS) development, including its influence on proliferation, neurogenesis, neuronal differentiation, cell death and synaptic plasticity [46,89-92]. These multiple biological functions of DYRK1A are due to its interactions with numerous cytoskeletal, synaptic and nuclear proteins, including transcription and splicing factors [93]. Together with other studies [85,94-96], these data strongly support the involvement of Dyrk1A in several neuropathological phenotypes and in the cognitive deficits that characterize Down syndrome. More recently, the observation that DYRK1A is overexpressed in the adult DS brain [97] implicated this protein also in the DS neurodegenerative phenotype. In particular, DYRK1A overexpression appears to be the cause of gene dosage-dependent modifications of several mechanisms that may contribute to the early onset of neurofibrillary degeneration. In fact, it has been demonstrated that Dyrk1A phosphorylates tau at several sites in vitro [98] and such sites are phosphorylated in DS brain [99]. Dyrk1A-induced tau phosphorylation inhibits the biological activity of tau, primes it for further phosphorylation by glycogen synthetase-3β (GSK- 3β) and promotes its self-aggregation into neurofibrillary tangles (NFTs) [99]. Interestingly, besides to phosphorylating protein, DYRK1A also colocalizes with NFTs [100]. In addition, neuropathological and molecular studies indicate that overexpressed nuclear DYRK1A contributes to the modification of the alternative splicing of Tau leading to neurofibrillary degeneration [101,102]. Neurofibrillary degeneration is the leading cause of neuronal death and dementia in Alzheimer’s disease (AD) and in DS⁄AD. The multi-pathway involvement of DYRK1A in neurofibrillary degeneration indicates that therapeutic inhibition of the activity of overexpressed DYRK1A may delay the age of onset and inhibit the progression of neurodegeneration in DS. To this regard, the studies recently performed by the group of Delabar [103] represent, arguably, the best example of how the functional knowledge about DSCR genes can be translated into new potential therapeutic strategy. Indeed, this research group has found that Epigallocatechin gallate (EGCG) - a member of a natural polyphenols family, found in great amount in green tea leaves - is a specific and safe DYRK1A inhibitor and that its administration can revert the brain defects induced by overexpression of DYRK1A [103]. Together with a previous report showing that EGCG administration may beneficially affect the LTP abnormalities detected in Ts65Dn mice [104], this study paved the way for the promotion of clinical trials, which are already in Phase 2 (see for instance http://clinicaltrials.gov/ct2/show/NCT01394796).

3.3.2. TTC3

Since its discovery in 1996, the TTC3 gene has been considered an important candidate for the CNS-related phenotypes that characterize DS, because of its mapping within the DSCR [105,106]. This hypothesis was further supported by the analysis of TTC3 expression during normal development. Indeed, during mouse and human brain embryogenesis, TTC3 expression shows regional and cellular specificities well correlated with the anatomical defects observed in DS patients [55,107]. In particular, TTC3 is expressed at highest levels in the post-mitotic areas of central nervous system (CNS), suggesting a role in neuronal cell differentiation [108,109]. Moreover, it has been reported that the expression of TTC3 is increased in tissues and in cells derived from DS experimental models [110] and from DS individuals [111,112]. In 2007, on the basis of both overexpression and knockdown experiments performed in PC12 neuroblastoma cells, we demonstrated that the TTC3 protein may play a pivotal role in regulating the differentiation program of neuronal cells, starting from the earliest stages [113]. More specifically, increased TTC3 function strongly prevents the neurite sprouting normally elicited by NGF-treatment, while TTC3 knockdown increases neurite length [113]. Importantly, TTC3 may affect not only the generation of neuronal processes, but also their maintenance (Berto et al., unpublished)., and its effects on neuronal differentiation are mediated by the activation of a specific pathway comprising the master cytoskeletal regulator RhoA and its effettor proteins, namely Citron-isoforms [113] Rho kinases (ROCKs) and LIM-kinase (Berto et al., in preparation), which have been implicated in all the different aspects of the neuronal differentiation program [114] and in different aspect of cognitive disorders [115]. Importantly, specific inhibitors of ROCKs, such as Fasudil, have been already approved by FDA, and therefore represent ideal candidates for testing in the experimental models [116]. In addition, a recent report by the group of Dr. M. Noguchi has shown that TTC3 can down-modulate the activity of the Akt kinases (AKTs), by promoting their ubiquitination and degradation [111]. This observation is particularly important, not only because AKTs have been shown to regulate neuronal survival [117], axonogenesis [118], dendritogenesis and synaptogenesis [119], but especially because these proteins are effectors of the PI3K pathway, which is the subject of extensive pharmacological investigation, in light of its centrality in cancer and inflammation research [120,121].

3.4. Gene networks affecting the excitatory-inhibitory balance in DS

The majority of forebrain is comprised of excitatory glutamatergic projection neurons and approximately 10% inhibitory γ-amminobutyric acid (GABA) interneurons. The normal functioning of the neural networks underlying cognitive functions depend on a finely-tuned balance of excitatory and inhibitory activities [122]. Accordingly, different reports have supported the possibility that cognitive impairment in DS models can be related to specific alterations of the excitatory/inhibitory balance, which may result from the direct action of DSCR genes or from more indirect mechanisms. For instance, it has been hypothesized that the increased dosage of HSA21 gene could favor the excitatory inputs in the hippocampus by increasing the activity of N-methyl-D-aspartate (NMDA) receptor (NMDAR), with potential effects on synaptic plasticity and neuron survival [123]. This theory was based on the observation that that several HSA21 genes, such as APP, SOD1, RCAN1 and DYRK1A, directly interact or indirectly affect the activity of the NMDARs. The best characterized pathway is that involving RCAN1, which regulates NMDARs by directly binding and inhibiting the calcineurin protein phosphatase (CaN) [71,77,124]. NMDARs are CaN targets [125] [126] and CaN inhibition leads to increased NMDARs [127] activity, by decreasing channel open probability and mean time [127]. On this basis Costa and co-workers hypothesized that the noncompetitive NMDA antagonist memantine, which acts as open channel blocker and is currently approved for AD therapy, could mimic the actions of CaN and restore normal NMDARs function, possibly improving learning and memory [123]. Indeed, memantine ameliorates contextual fear conditioning learning in 4–6- and 10–14-month old Ts65Dn mice when administered at 5 mg/kg by acute intraperitoneal injection before context exposure. Despite these studies, a recently published clinical trial reported that memantine is not an effective pharmacological treatment for cognitive decline or dementia in DS patients who are above 40 years old [128]. This suggests that therapies that are effective in DS models and in AD patients may not necessarily confer benefits in DS.

More consistent reports have shown that the LTP phenotypes and the reduced performance in cognitive tests observed in mouse models could be the result of excessive GABA-ergic responses, producing a net decrease of synaptic output [36,37,129]. This phenomenon could be a direct effect of the overexpression of at least three proteins encoded by the DSCR, namely the chloride channel CLIC6 and the rectifying potassium channels KCNJ6 and KCNJ15. Accordingly, primary hippocampal neurons derived from Ts65Dn mice display a significant increase in GABA-mediated GIRK currents, consistent with the increased expression of KCNJ6/GIRK2 [130]. However, some of the data are also consistent with an increased pre-synaptic availability of GABA [129], produced by undefined and probably indirect mechanisms. On this basis, several pharmacological interventions have been proposed to restore the excitatory-inhibitory imbalance by decreasing the excessive inhibition of GABAergic neurotransmission prevalent in DS mouse models [131]. In particular, Ts65Dn mice have been treated with non-competitive GABAA antagonists, pentylenetetrazol (PTZ) and picrotoxin (PTX), which inhibit GABAA receptors. Chronic treatment with PTZ reversed the deficits seen in the novel object recognition task (NORT) and spontaneous alternation tasks in Ts65Dn mice [129,132]. Surprisingly, the improvement in cognition and LTP was sustained for up to 2 months after initial treatment, suggesting a long-lasting effect on neuronal circuit modification. Chronic treatment with PTZ for 8 weeks in Ts65Dn mice did not modify sensorimotor abilities and locomotor activity in home cages. However it did rescue learning and memory performance in the Morris water maze (MWM) task [133]. Recently, chronic treatment in Ts65Dn mice with an inverse agonist selective for the α5 subunit of the GABAA benzodiazepine receptor (α5IA) improved cognitive deficits in the MWM and normalized Sod1 overexpression with an enhancement in learning-evoked immediate early genes expression levels [134]. Encouraged by this body of evidence, Roche, a healthcare company, recently announced the commencement of a trial to examine the cognitive impact of reducing GABA-ergic neurotransmission in the hippocampus using a drug selective for the α5 subunit of GABAA receptors (http://www.roche-trials.com).

Finally, the imbalance in excitatory/inhibitory ratio could be the result of abnormal neurogenesis. Indeed, reduced cell numbers in the DS hippocampus could be caused by impaired adult neurogenesis, which has been observed in Ts65Dn [135] [136] and Ts1Cje mice [137]. Therefore, approaches targeting neurogenesis seem very promising for DS therapy. Interestingly, a fascinating connection has been documented between the DSCR gene KCNJ6 and adult neurogenesis, mediated by serotonin signaling. DS has long been associated with defects in the serotonergic system [138]. In particular, the serotonin 5-HT1A receptor expression peaks earlier in developing DS brains and decreases to below normal levels by birth [139]. Moreover reduced 5-HT levels are present in adults with DS [140]. Since 5-HT depletion causes a permanent reduction in neuron number in the adult brain [138], it is conceivable that alterations in the serotonergic systems during early life stages may contribute to the reduced neurogenesis of the DS brain. Activity of the serotonin receptor 1A (5HTR1A) is required for adult neurogenesis in the hippocampus [141] and is mediated by the potassium channel KCNJ6. Overexpression of KCNJ6, as in the Ts65Dn, may over-inhibit presynaptic 5HTR1A, causing reduced levels of serotonin. Fluoxetine, an antidepressant that inhibits serotonin (5-HT) reuptake, inhibits KCNJ6 and increases presynaptic levels of serotonin. Consistent with this, it has been already demonstrated that fluoxetine is able to rescue neurogenesis in the adult Ts65Dn [135]. Recently, treatment during the early postnatal period restored neurogenesis and the total number of neurons in the dentate gyrus. This effect was accompanied by the full recovery of a cognitive task [142]. The releance of these data is even greater if considering that fluoxetin is an antidepressant widely used by adults and prescribed in children and adolescents [143] and that it does not seem to have negative effects on post-natal development [144].

3.5. The DSCR and Alzheimer-related molecular networks

Most DS patients experience a decline in cognition during adulthood, followed by the development of classical Alzheimer’s disease (AD) neuropathology, characterized by the accumulation of amyloid plaques containing high levels of the A-beta fragments of the APP protein, by neurofibrillary tangles containing high levels of hyperphosphorylated Tau protein and by massive neurodegeneration [145]. Increased dosage of the APP gene, which is located outside the DSCR, is very likely the most important factor that underlies this phenomenon [146]. Indeed, increased dosage of APP is sufficient to strongly increase the risk of AD, since APP gene duplication has been detected as the mutation responsible for some early-onset familial cases of AD [147]. The link between AD and the APP gene has been further strengthened by the finding that an extra copy of APP seems to be necessary for the development of AD in DS. Indeed, it has been reported the case of an old patient affected by DS but not showing any signs of dementia [148]. At autopsy, plaques and tangles were absent in the brain of this individual. The patient had a segmental trisomy HSA21, not including the APP gene [148]. These data strongly support that the early onset of AD pathology in DS is in part due to overexpression of the APP gene. The data obtained from experimental models further support the crucial role of APP in DS [51]. Indeed, it has been shown that APP overexpression in Ts65Dn impairs the retrograde transport of nerve growth factor (NGF) from the hippocampus to the basal forebrain, causing the degeneration of BFCN [51], which significantly degenerates in Ts65Dn. Importantly, APP is one of the few genes for which a successful subtractive genetic approach has been reported, since restoring APP gene dosage to two copies in the Ts65Dn model corrected the water maze phenotype and prevented BFCN degeneration [50,51]. Finally, APP-mediated pathological mechanism may also contribute to the developmental abnormalities detected in mouse models, since it has been suggested that APP overexpression can result in increased Notch signaling pathway, which is crucial for neuronal and glial differentiation [149]. However, it is conceivable that also some of the DSCR genes may cooperate with APP in accelerating the AD-related neuropathological phenotypes observed in DS patients. In particular BACE2 could promote the beta-cleavage of APP, further increasing the amount of generated A-beta peptides [150-152]. DYRK1A can also play an important role, because it can stimulate the phosphorylation of APP and Tau, resulting in increased cleavage and aggregation, respectively [98,153]. Finally, Tau hyperphosphorylation can be stimulated by increased expression of RCAN1, since phosphorylated Tau is one of the substrates of calcineurin [154]. Moreover, it has been shown that this activity of RCAN1 can be modulated by DYRK1A [155] Therefore it is very likely that the development of new approaches aimed at targeting these proteins could turn out to be beneficial both for AD and for DS management.

3.6. DSCR-dependent RNA-networks

As it is generally the case for the human genome, besides to protein coding genes, the DSCR contains many sequences that have been so far almost completely neglected, because they are not predicted to encode for proteins [16]. However, as we show in Table 2, on the basis of the current knowledge, many of these loci display features indicating that they could be functionally relevant and could contribute to the pathogenesis of DS phenotypes. Indeed, besides to the two copies of snRNAs and five copies of snoRNAs associated to splicing factors, the DSCR contains many regions that are transcribed to produce processed transcripts, devoid of coding potential. Some of these sequences, such as antisense transcripts, processed pseudogenes and sequences located in proximity of promoters, are closely associated to functioning genes, and could be involved in their regulation, as it has been shown in many other cases [156-158]. In many other cases, the genes appear to produce llincRNAs, that could act in cis to modify chromatin structure, or in trans to modify gene expression at the transcriptional and post transcriptional level, as it has been shown in the cases of HOTAIR [159] and of LincRNA-p21 [160,161]. Although the function of these molecules is at the moment completely unknown, their study could be extremely interesting. Indeed many of these sequences have been implicated in the epigenetic and in the post-transcriptional control of gene expression. Moreover, since these sequences diverge much more rapidly than the sequences of protein-coding genes, it is very likely that they could be strongly implicated in the control of human-specific features and phenotypes. Therefore, it seems reasonable to anticipate that the functional study of lincRNA-encoding genes in DS models and the study of their variation in humans will be a fertile ground for future research. Finally, the DSCR contains at least three genes encoding miRNA precursors (probably five, if considering also those that have only been predicted). Interestingly, mir-802, which is encoded by the DSCR, and mir-155, which is located on HSA21 in a more centromeric position, have been shown to repress the expression of MeCP2 [162], whose inactivation is the cause of Rett syndrome. Since MeCP2 is also repressed by HMGN1, this study further underscore the potential relevance of MeCP2 repression in DS and provides a very interesting example of how the intertwining of transcription and post-transcriptional regulatory networks dependent on DSCR genes can produce intellectual disability. Considering the reported reversibility of MeCP2 downregulation phenotypes [163] and the great efforts that are being dedicated to identify drugable pathways downstream of MeCP2 [164], it is conceivable that the functional exploration of these networks in DS could be also relevant for the development of future therapies.

Advertisement

4. Concluding remarks

Functional information on HSA21 genes is still quite partial and mostly limited to a subset of protein-coding genes. However, the recent success in DS models of therapeutic strategies targeted either on specific DSCR genes, or even on much broader mechanisms, justifies to our opinion an optimistic view of the future. In particular, we think that it will be reasonable to expect that a high level of understanding of the complex networks implicating DSCR genes through systems biology approaches will provide very useful insight, which could be translated into new therapies that could turn out to be useful not only for DS, but also for other disorders such as Alzheimer’s disease and Rett syndrome.

Advertisement

Acknowledgments

We are grateful to Dr. Christian Damasco for his help in the production of Tables 1 and 2. The financial contribution of the Jerome Lejeune Foundation FDC and GB is gratefully acknowledged.

References

  1. 1. Piro RM (2012) Network medicine: linking disorders. Hum Genet.
  2. 2. Chan SY, Loscalzo J (2012) The emerging paradigm of network medicine in the study of human disease. Circ Res 111: 359-374.
  3. 3. Jackson JF, North ER, 3rd, Thomas JG (1976) Clinical diagnosis of Down's syndrome. Clin Genet 9: 483-487.
  4. 4. Antonarakis SE, Epstein CJ (2006) The challenge of Down syndrome. Trends Mol Med 12: 473-479.
  5. 5. Lejeune J, Gautier M, Turpin R (1959) Study of somatic chromosomes from 9 mongoloid children. C R Hebd Seances Acad Sci 248: 1721-1722.
  6. 6. Antonarakis SE (1998) 10 years of Genomics, chromosome 21, and Down syndrome. Genomics 51: 1-16.
  7. 7. Poissonnier M, Saint-Paul B, Dutrillaux B, Chassaigne M, Gruyer P, et al. (1976) Partial trisomy 21 (21q21 - 21q22.2). Ann Genet 19: 69-73.
  8. 8. Dierssen M, Herault Y, Estivill X (2009) Aneuploidy: from a physiological mechanism of variance to Down syndrome. Physiol Rev 89: 887-920.
  9. 9. Rachidi M, Lopes C, Vayssettes C, Smith DJ, Rubin EM, et al. (2007) New cerebellar phenotypes in YAC transgenic mouse in vivo library of human Down syndrome critical region-1. Biochem Biophys Res Commun 364: 488-494.
  10. 10. Delabar JM, Theophile D, Rahmani Z, Chettouh Z, Blouin JL, et al. (1993) Molecular mapping of twenty-four features of Down syndrome on chromosome 21. Eur J Hum Genet 1: 114-124.
  11. 11. Korenberg JR, Chen XN, Schipper R, Sun Z, Gonsky R, et al. (1994) Down syndrome phenotypes: the consequences of chromosomal imbalance. Proc Natl Acad Sci U S A 91: 4997-5001.
  12. 12. Chabert C, Cherfouh A, Delabar JM, Duquenne V (2001) Assessing implications between genotypic and phenotypic variables through lattice analysis. Behav Genet 31: 125-139.
  13. 13. Lyle R, Bena F, Gagos S, Gehrig C, Lopez G, et al. (2009) Genotype-phenotype correlations in Down syndrome identified by array CGH in 30 cases of partial trisomy and partial monosomy chromosome 21. Eur J Hum Genet 17: 454-466.
  14. 14. Kahlem P, Sultan M, Herwig R, Steinfath M, Balzereit D, et al. (2004) Transcript level alterations reflect gene dosage effects across multiple tissues in a mouse model of down syndrome. Genome Res 14: 1258-1267.
  15. 15. Hattori M, Fujiyama A, Taylor TD, Watanabe H, Yada T, et al. (2000) The DNA sequence of human chromosome 21. Nature 405: 311-319.
  16. 16. Esteller M (2011) Non-coding RNAs in human disease. Nat Rev Genet 12: 861-874.
  17. 17. Tay Y, Kats L, Salmena L, Weiss D, Tan SM, et al. (2011) Coding-independent regulation of the tumor suppressor PTEN by competing endogenous mRNAs. Cell 147: 344-357.
  18. 18. Sumazin P, Yang X, Chiu HS, Chung WJ, Iyer A, et al. (2011) An extensive microRNA-mediated network of RNA-RNA interactions regulates established oncogenic pathways in glioblastoma. Cell 147: 370-381.
  19. 19. Salmena L, Poliseno L, Tay Y, Kats L, Pandolfi PP (2011) A ceRNA hypothesis: the Rosetta Stone of a hidden RNA language? Cell 146: 353-358.
  20. 20. Cesana M, Cacchiarelli D, Legnini I, Santini T, Sthandier O, et al. (2011) A long noncoding RNA controls muscle differentiation by functioning as a competing endogenous RNA. Cell 147: 358-369.
  21. 21. Poliseno L, Salmena L, Zhang J, Carver B, Haveman WJ, et al. (2011) A coding-independent function of gene and pseudogene mRNAs regulates tumour biology. Nature 465: 1033-1038.
  22. 22. Davisson MT, Gardiner K, Costa AC (2001) Report and abstracts of the ninth international workshop on the molecular biology of human chromosome 21 and Down syndrome. Bar Harbor, Maine, USA. 23-26 September 2000. Cytogenet Cell Genet 92: 1-22.
  23. 23. Davisson MT, Schmidt C, Akeson EC (1990) Segmental trisomy of murine chromosome 16: a new model system for studying Down syndrome. Prog Clin Biol Res 360: 263-280.
  24. 24. Sago H, Carlson EJ, Smith DJ, Kilbridge J, Rubin EM, et al. (1998) Ts1Cje, a partial trisomy 16 mouse model for Down syndrome, exhibits learning and behavioral abnormalities. Proc Natl Acad Sci U S A 95: 6256-6261.
  25. 25. Sago H, Carlson EJ, Smith DJ, Rubin EM, Crnic LS, et al. (2000) Genetic dissection of region associated with behavioral abnormalities in mouse models for Down syndrome. Pediatr Res 48: 606-613.
  26. 26. Demas GE, Nelson RJ, Krueger BK, Yarowsky PJ (1998) Impaired spatial working and reference memory in segmental trisomy (Ts65Dn) mice. Behav Brain Res 90: 199-201.
  27. 27. Demas GE, Nelson RJ, Krueger BK, Yarowsky PJ (1996) Spatial memory deficits in segmental trisomic Ts65Dn mice. Behav Brain Res 82: 85-92.
  28. 28. Holtzman DM, Santucci D, Kilbridge J, Chua-Couzens J, Fontana DJ, et al. (1996) Developmental abnormalities and age-related neurodegeneration in a mouse model of Down syndrome. Proc Natl Acad Sci U S A 93: 13333-13338.
  29. 29. Hyde LA, Crnic LS (2001) Age-related deficits in context discrimination learning in Ts65Dn mice that model Down syndrome and Alzheimer's disease. Behav Neurosci 115: 1239-1246.
  30. 30. Escorihuela RM, Fernandez-Teruel A, Vallina IF, Baamonde C, Lumbreras MA, et al. (1995) A behavioral assessment of Ts65Dn mice: a putative Down syndrome model. Neurosci Lett 199: 143-146.
  31. 31. Escorihuela RM, Vallina IF, Martinez-Cue C, Baamonde C, Dierssen M, et al. (1998) Impaired short- and long-term memory in Ts65Dn mice, a model for Down syndrome. Neurosci Lett 247: 171-174.
  32. 32. Reeves RH, Irving NG, Moran TH, Wohn A, Kitt C, et al. (1995) A mouse model for Down syndrome exhibits learning and behaviour deficits. Nat Genet 11: 177-184.
  33. 33. Belichenko PV, Masliah E, Kleschevnikov AM, Villar AJ, Epstein CJ, et al. (2004) Synaptic structural abnormalities in the Ts65Dn mouse model of Down Syndrome. J Comp Neurol 480: 281-298.
  34. 34. Siarey RJ, Stoll J, Rapoport SI, Galdzicki Z (1997) Altered long-term potentiation in the young and old Ts65Dn mouse, a model for Down Syndrome. Neuropharmacology 36: 1549-1554.
  35. 35. Siarey RJ, Carlson EJ, Epstein CJ, Balbo A, Rapoport SI, et al. (1999) Increased synaptic depression in the Ts65Dn mouse, a model for mental retardation in Down syndrome. Neuropharmacology 38: 1917-1920.
  36. 36. Kleschevnikov AM, Belichenko PV, Villar AJ, Epstein CJ, Malenka RC, et al. (2004) Hippocampal long-term potentiation suppressed by increased inhibition in the Ts65Dn mouse, a genetic model of Down syndrome. J Neurosci 24: 8153-8160.
  37. 37. Siarey RJ, Villar AJ, Epstein CJ, Galdzicki Z (2005) Abnormal synaptic plasticity in the Ts1Cje segmental trisomy 16 mouse model of Down syndrome. Neuropharmacology 49: 122-128.
  38. 38. Belichenko PV, Kleschevnikov AM, Salehi A, Epstein CJ, Mobley WC (2007) Synaptic and cognitive abnormalities in mouse models of Down syndrome: exploring genotype-phenotype relationships. J Comp Neurol 504: 329-345.
  39. 39. Olson LE, Richtsmeier JT, Leszl J, Reeves RH (2004) A chromosome 21 critical region does not cause specific Down syndrome phenotypes. Science 306: 687-690.
  40. 40. Aldridge K, Reeves RH, Olson LE, Richtsmeier JT (2007) Differential effects of trisomy on brain shape and volume in related aneuploid mouse models. Am J Med Genet A 143A: 1060-1070.
  41. 41. Olson LE, Roper RJ, Sengstaken CL, Peterson EA, Aquino V, et al. (2007) Trisomy for the Down syndrome 'critical region' is necessary but not sufficient for brain phenotypes of trisomic mice. Hum Mol Genet 16: 774-782.
  42. 42. Belichenko NP, Belichenko PV, Kleschevnikov AM, Salehi A, Reeves RH, et al. (2009) The "Down syndrome critical region" is sufficient in the mouse model to confer behavioral, neurophysiological, and synaptic phenotypes characteristic of Down syndrome. J Neurosci 29: 5938-5948.
  43. 43. Smith DJ, Stevens ME, Sudanagunta SP, Bronson RT, Makhinson M, et al. (1997) Functional screening of 2 Mb of human chromosome 21q22.2 in transgenic mice implicates minibrain in learning defects associated with Down syndrome. Nat Genet 16: 28-36.
  44. 44. Smith DJ, Rubin EM (1997) Functional screening and complex traits: human 21q22.2 sequences affecting learning in mice. Hum Mol Genet 6: 1729-1733.
  45. 45. Smith DJ, Zhu Y, Zhang J, Cheng JF, Rubin EM (1995) Construction of a panel of transgenic mice containing a contiguous 2-Mb set of YAC/P1 clones from human chromosome 21q22.2. Genomics 27: 425-434.
  46. 46. Ahn KJ, Jeong HK, Choi HS, Ryoo SR, Kim YJ, et al. (2006) DYRK1A BAC transgenic mice show altered synaptic plasticity with learning and memory defects. Neurobiol Dis 22: 463-472.
  47. 47. Liu C, Belichenko PV, Zhang L, Fu D, Kleschevnikov AM, et al. (2011) Mouse models for Down syndrome-associated developmental cognitive disabilities. Dev Neurosci 33: 404-413.
  48. 48. Baek KH, Zaslavsky A, Lynch RC, Britt C, Okada Y, et al. (2009) Down's syndrome suppression of tumour growth and the role of the calcineurin inhibitor DSCR1. Nature 459: 1126-1130.
  49. 49. Chakrabarti L, Best TK, Cramer NP, Carney RS, Isaac JT, et al. (2010) Olig1 and Olig2 triplication causes developmental brain defects in Down syndrome. Nat Neurosci 13: 927-934.
  50. 50. Cataldo AM, Petanceska S, Peterhoff CM, Terio NB, Epstein CJ, et al. (2003) App gene dosage modulates endosomal abnormalities of Alzheimer's disease in a segmental trisomy 16 mouse model of down syndrome. J Neurosci 23: 6788-6792.
  51. 51. Salehi A, Delcroix JD, Belichenko PV, Zhan K, Wu C, et al. (2006) Increased App expression in a mouse model of Down's syndrome disrupts NGF transport and causes cholinergic neuron degeneration. Neuron 51: 29-42.
  52. 52. Haydar TF, Reeves RH (2012) Trisomy 21 and early brain development. Trends Neurosci 35: 81-91.
  53. 53. Antonarakis SE, Lyle R, Dermitzakis ET, Reymond A, Deutsch S (2004) Chromosome 21 and down syndrome: from genomics to pathophysiology. Nat Rev Genet 5: 725-738.
  54. 54. Kleschevnikov AM, Belichenko PV, Faizi M, Jacobs LF, Htun K, et al. (2012) Deficits in Cognition and Synaptic Plasticity in a Mouse Model of Down Syndrome Ameliorated by GABAB Receptor Antagonists. J Neurosci 32: 9217-9227.
  55. 55. Golden JA, Hyman BT (1994) Development of the superior temporal neocortex is anomalous in trisomy 21. J Neuropathol Exp Neurol 53: 513-520.
  56. 56. Guidi S, Bonasoni P, Ceccarelli C, Santini D, Gualtieri F, et al. (2008) Neurogenesis impairment and increased cell death reduce total neuron number in the hippocampal region of fetuses with Down syndrome. Brain Pathol 18: 180-197.
  57. 57. Vuksic M, Petanjek Z, Rasin MR, Kostovic I (2002) Perinatal growth of prefrontal layer III pyramids in Down syndrome. Pediatr Neurol 27: 36-38.
  58. 58. Becker L, Mito T, Takashima S, Onodera K (1991) Growth and development of the brain in Down syndrome. Prog Clin Biol Res 373: 133-152.
  59. 59. Takashima S, Becker LE, Armstrong DL, Chan F (1981) Abnormal neuronal development in the visual cortex of the human fetus and infant with down's syndrome. A quantitative and qualitative Golgi study. Brain Res 225: 1-21.
  60. 60. Becker LE, Armstrong DL, Chan F (1986) Dendritic atrophy in children with Down's syndrome. Ann Neurol 20: 520-526.
  61. 61. Hollenhorst PC, McIntosh LP, Graves BJ (2011) Genomic and biochemical insights into the specificity of ETS transcription factors. Annu Rev Biochem 80: 437-471.
  62. 62. Dryden NH, Sperone A, Martin-Almedina S, Hannah RL, Birdsey GM, et al. (2012) The transcription factor Erg controls endothelial cell quiescence by repressing activity of nuclear factor (NF)-kappaB p65. J Biol Chem 287: 12331-12342.
  63. 63. Martens JH (2011) Acute myeloid leukemia: a central role for the ETS factor ERG. Int J Biochem Cell Biol 43: 1413-1416.
  64. 64. Fisher IB, Ostrowski M, Muthusamy N (2012) Role for Ets-2(Thr-72) transcription factor in stage-specific thymocyte development and survival. J Biol Chem 287: 5199-5210.
  65. 65. Lam K, Zhang DE (2012) RUNX1 and RUNX1-ETO: roles in hematopoiesis and leukemogenesis. Front Biosci 17: 1120-1139.
  66. 66. Huang H, Rambaldi I, Daniels E, Featherstone M (2003) Expression of the Wdr9 gene and protein products during mouse development. Dev Dyn 227: 608-614.
  67. 67. Abuhatzira L, Shamir A, Schones DE, Schaffer AA, Bustin M (2011) The chromatin-binding protein HMGN1 regulates the expression of methyl CpG-binding protein 2 (MECP2) and affects the behavior of mice. J Biol Chem 286: 42051-42062.
  68. 68. Martin-Bermudo MD, Carmena A, Jimenez F (1995) Neurogenic genes control gene expression at the transcriptional level in early neurogenesis and in mesectoderm specification. Development 121: 219-224.
  69. 69. Havis E, Coumailleau P, Bonnet A, Bismuth K, Bonnin MA, et al. (2012) Sim2 prevents entry into the myogenic program by repressing MyoD transcription during limb embryonic myogenesis. Development 139: 1910-1920.
  70. 70. Fuentes JJ, Pritchard MA, Planas AM, Bosch A, Ferrer I, et al. (1995) A new human gene from the Down syndrome critical region encodes a proline-rich protein highly expressed in fetal brain and heart. Hum Mol Genet 4: 1935-1944.
  71. 71. Fuentes JJ, Genesca L, Kingsbury TJ, Cunningham KW, Perez-Riba M, et al. (2000) DSCR1, overexpressed in Down syndrome, is an inhibitor of calcineurin-mediated signaling pathways. Hum Mol Genet 9: 1681-1690.
  72. 72. Rothermel B, Vega RB, Yang J, Wu H, Bassel-Duby R, et al. (2000) A protein encoded within the Down syndrome critical region is enriched in striated muscles and inhibits calcineurin signaling. J Biol Chem 275: 8719-8725.
  73. 73. Ermak G, Morgan TE, Davies KJ (2001) Chronic overexpression of the calcineurin inhibitory gene DSCR1 (Adapt78) is associated with Alzheimer's disease. J Biol Chem 276: 38787-38794.
  74. 74. Chang KT, Min KT (2005) Drosophila melanogaster homolog of Down syndrome critical region 1 is critical for mitochondrial function. Nat Neurosci 8: 1577-1585.
  75. 75. Ermak G, Cheadle C, Becker KG, Harris CD, Davies KJ (2004) DSCR1(Adapt78) modulates expression of SOD1. Faseb J 18: 62-69.
  76. 76. Gulesserian T, Seidl R, Hardmeier R, Cairns N, Lubec G (2001) Superoxide dismutase SOD1, encoded on chromosome 21, but not SOD2 is overexpressed in brains of patients with Down syndrome. J Investig Med 49: 41-46.
  77. 77. Ermak G, Harris CD, Davies KJ (2002) The DSCR1 (Adapt78) isoform 1 protein calcipressin 1 inhibits calcineurin and protects against acute calcium-mediated stress damage, including transient oxidative stress. Faseb J 16: 814-824.
  78. 78. Jovanovic SV, Clements D, MacLeod K (1998) Biomarkers of oxidative stress are significantly elevated in Down syndrome. Free Radic Biol Med 25: 1044-1048.
  79. 79. Pallardo FV, Degan P, d'Ischia M, Kelly FJ, Zatterale A, et al. (2006) Multiple evidence for an early age pro-oxidant state in Down Syndrome patients. Biogerontology 7: 211-220.
  80. 80. Okui M, Ide T, Morita K, Funakoshi E, Ito F, et al. (1999) High-level expression of the Mnb/Dyrk1A gene in brain and heart during rat early development. Genomics 62: 165-171.
  81. 81. Tejedor F, Zhu XR, Kaltenbach E, Ackermann A, Baumann A, et al. (1995) minibrain: a new protein kinase family involved in postembryonic neurogenesis in Drosophila. Neuron 14: 287-301.
  82. 82. Chen H, Antonarakis SE (1997) Localisation of a human homologue of the Drosophila mnb and rat Dyrk genes to chromosome 21q22.2. Hum Genet 99: 262-265.
  83. 83. Fotaki V, Dierssen M, Alcantara S, Martinez S, Marti E, et al. (2002) Dyrk1A haploinsufficiency affects viability and causes developmental delay and abnormal brain morphology in mice. Mol Cell Biol 22: 6636-6647.
  84. 84. Moller RS, Kubart S, Hoeltzenbein M, Heye B, Vogel I, et al. (2008) Truncation of the Down syndrome candidate gene DYRK1A in two unrelated patients with microcephaly. Am J Hum Genet 82: 1165-1170.
  85. 85. Altafaj X, Dierssen M, Baamonde C, Marti E, Visa J, et al. (2001) Neurodevelopmental delay, motor abnormalities and cognitive deficits in transgenic mice overexpressing Dyrk1A (minibrain), a murine model of Down's syndrome. Hum Mol Genet 10: 1915-1923.
  86. 86. Rahmani Z, Lopes C, Rachidi M, Delabar JM (1998) Expression of the mnb (dyrk) protein in adult and embryonic mouse tissues. Biochem Biophys Res Commun 253: 514-518.
  87. 87. Guimera J, Casas C, Estivill X, Pritchard M (1999) Human minibrain homologue (MNBH/DYRK1): characterization, alternative splicing, differential tissue expression, and overexpression in Down syndrome. Genomics 57: 407-418.
  88. 88. Lyle R, Gehrig C, Neergaard-Henrichsen C, Deutsch S, Antonarakis SE (2004) Gene expression from the aneuploid chromosome in a trisomy mouse model of down syndrome. Genome Res 14: 1268-1274.
  89. 89. Park J, Oh Y, Yoo L, Jung MS, Song WJ, et al. (2010) Dyrk1A phosphorylates p53 and inhibits proliferation of embryonic neuronal cells. J Biol Chem 285: 31895-31906.
  90. 90. Hammerle B, Elizalde C, Tejedor FJ (2008) The spatio-temporal and subcellular expression of the candidate Down syndrome gene Mnb/Dyrk1A in the developing mouse brain suggests distinct sequential roles in neuronal development. Eur J Neurosci 27: 1061-1074.
  91. 91. Hammerle B, Ulin E, Guimera J, Becker W, Guillemot F, et al. (2011) Transient expression of Mnb/Dyrk1a couples cell cycle exit and differentiation of neuronal precursors by inducing p27KIP1 expression and suppressing NOTCH signaling. Development 138: 2543-2554.
  92. 92. Yabut O, Domogauer J, D'Arcangelo G (2010) Dyrk1A overexpression inhibits proliferation and induces premature neuronal differentiation of neural progenitor cells. J Neurosci 30: 4004-4014.
  93. 93. Galceran J, de Graaf K, Tejedor FJ, Becker W (2003) The MNB/DYRK1A protein kinase: genetic and biochemical properties. J Neural Transm Suppl: 139-148.
  94. 94. Hammerle B, Elizalde C, Galceran J, Becker W, Tejedor FJ (2003) The MNB/DYRK1A protein kinase: neurobiological functions and Down syndrome implications. J Neural Transm Suppl: 129-137.
  95. 95. Hammerle B, Carnicero A, Elizalde C, Ceron J, Martinez S, et al. (2003) Expression patterns and subcellular localization of the Down syndrome candidate protein MNB/DYRK1A suggest a role in late neuronal differentiation. Eur J Neurosci 17: 2277-2286.
  96. 96. Martinez de Lagran M, Altafaj X, Gallego X, Marti E, Estivill X, et al. (2004) Motor phenotypic alterations in TgDyrk1a transgenic mice implicate DYRK1A in Down syndrome motor dysfunction. Neurobiol Dis 15: 132-142.
  97. 97. Dowjat WK, Adayev T, Kuchna I, Nowicki K, Palminiello S, et al. (2007) Trisomy-driven overexpression of DYRK1A kinase in the brain of subjects with Down syndrome. Neurosci Lett 413: 77-81.
  98. 98. Ryoo SR, Jeong HK, Radnaabazar C, Yoo JJ, Cho HJ, et al. (2007) DYRK1A-mediated hyperphosphorylation of Tau. A functional link between Down syndrome and Alzheimer disease. J Biol Chem 282: 34850-34857.
  99. 99. Liu F, Liang Z, Wegiel J, Hwang YW, Iqbal K, et al. (2008) Overexpression of Dyrk1A contributes to neurofibrillary degeneration in Down syndrome. Faseb J 22: 3224-3233.
  100. 100. Wegiel J, Dowjat K, Kaczmarski W, Kuchna I, Nowicki K, et al. (2008) The role of overexpressed DYRK1A protein in the early onset of neurofibrillary degeneration in Down syndrome. Acta Neuropathol 116: 391-407.
  101. 101. Shi J, Zhang T, Zhou C, Chohan MO, Gu X, et al. (2008) Increased dosage of Dyrk1A alters alternative splicing factor (ASF)-regulated alternative splicing of tau in Down syndrome. J Biol Chem 283: 28660-28669.
  102. 102. Yin X, Jin N, Gu J, Shi J, Gong CX, et al. (2012) Dual-specificity-tyrosine-phosphorylated and regulated kinase 1A (Dyrk1A ) modulates serine-arginine rich protein 55 (SRp55)-promoted tau exon 10 inclusion. J Biol Chem.
  103. 103. Guedj F, Sebrie C, Rivals I, Ledru A, Paly E, et al. (2009) Green tea polyphenols rescue of brain defects induced by overexpression of DYRK1A. PLoS One 4: e4606.
  104. 104. Xie W, Ramakrishna N, Wieraszko A, Hwang YW (2008) Promotion of neuronal plasticity by (-)-epigallocatechin-3-gallate. Neurochem Res 33: 776-783.
  105. 105. Ohira M, Ootsuyama A, Suzuki E, Ichikawa H, Seki N, et al. (1996) Identification of a novel human gene containing the tetratricopeptide repeat domain from the Down syndrome region of chromosome 21. DNA Res 3: 9-16.
  106. 106. Tsukahara F, Urakawa I, Hattori M, Hirai M, Ohba K, et al. (1998) Molecular characterization of the mouse mtprd gene, a homologue of human TPRD: unique gene expression suggesting its critical role in the pathophysiology of Down syndrome. J Biochem 123: 1055-1063.
  107. 107. Raz N, Torres IJ, Briggs SD, Spencer WD, Thornton AE, et al. (1995) Selective neuroanatomic abnormalities in Down's syndrome and their cognitive correlates: evidence from MRI morphometry. Neurology 45: 356-366.
  108. 108. Lopes C, Rachidi M, Gassanova S, Sinet PM, Delabar JM (1999) Developmentally regulated expression of mtprd, the murine ortholog of tprd, a gene from the Down syndrome chromosomal region 1. Mech Dev 84: 189-193.
  109. 109. Rachidi M, Lopes C, Gassanova S, Sinet PM, Vekemans M, et al. (2000) Regional and cellular specificity of the expression of TPRD, the tetratricopeptide Down syndrome gene, during human embryonic development. Mech Dev 93: 189-193.
  110. 110. Saran NG, Pletcher MT, Natale JE, Cheng Y, Reeves RH (2003) Global disruption of the cerebellar transcriptome in a Down syndrome mouse model. Hum Mol Genet 12: 2013-2019.
  111. 111. Suizu F, Hiramuki Y, Okumura F, Matsuda M, Okumura AJ, et al. (2009) The E3 ligase TTC3 facilitates ubiquitination and degradation of phosphorylated Akt. Dev Cell 17: 800-810.
  112. 112. Toker A (2009) TTC3 ubiquitination terminates Akt-ivation. Dev Cell 17: 752-754.
  113. 113. Berto G, Camera P, Fusco C, Imarisio S, Ambrogio C, et al. (2007) The Down syndrome critical region protein TTC3 inhibits neuronal differentiation via RhoA and Citron kinase. J Cell Sci 120: 1859-1867.
  114. 114. Govek EE, Newey SE, Van Aelst L (2005) The role of the Rho GTPases in neuronal development. Genes Dev 19: 1-49.
  115. 115. Newey SE, Velamoor V, Govek EE, Van Aelst L (2005) Rho GTPases, dendritic structure, and mental retardation. J Neurobiol 64: 58-74.
  116. 116. Gupta V, Ahsan F (2010) Inhalational therapy for pulmonary arterial hypertension: current status and future prospects. Crit Rev Ther Drug Carrier Syst 27: 313-370.
  117. 117. Dudek H, Datta SR, Franke TF, Birnbaum MJ, Yao R, et al. (1997) Regulation of neuronal survival by the serine-threonine protein kinase Akt. Science 275: 661-665.
  118. 118. Grider MH, Park D, Spencer DM, Shine HD (2009) Lipid raft-targeted Akt promotes axonal branching and growth cone expansion via mTOR and Rac1, respectively. J Neurosci Res 87: 3033-3042.
  119. 119. Majumdar D, Nebhan CA, Hu L, Anderson B, Webb DJ (2011) An APPL1/Akt signaling complex regulates dendritic spine and synapse formation in hippocampal neurons. Mol Cell Neurosci 46: 633-644.
  120. 120. Harris SJ, Foster JG, Ward SG (2009) PI3K isoforms as drug targets in inflammatory diseases: lessons from pharmacological and genetic strategies. Curr Opin Investig Drugs 10: 1151-1162.
  121. 121. Courtney KD, Corcoran RB, Engelman JA (2010) The PI3K pathway as drug target in human cancer. J Clin Oncol 28: 1075-1083.
  122. 122. Castillo PE, Chiu CQ, Carroll RC (2011) Long-term plasticity at inhibitory synapses. Curr Opin Neurobiol 21: 328-338.
  123. 123. Costa AC, Scott-McKean JJ, Stasko MR (2008) Acute injections of the NMDA receptor antagonist memantine rescue performance deficits of the Ts65Dn mouse model of Down syndrome on a fear conditioning test. Neuropsychopharmacology 33: 1624-1632.
  124. 124. Chang KT, Min KT (2009) Upregulation of three Drosophila homologs of human chromosome 21 genes alters synaptic function: implications for Down syndrome. Proc Natl Acad Sci U S A 106: 17117-17122.
  125. 125. Krupp JJ, Vissel B, Thomas CG, Heinemann SF, Westbrook GL (2002) Calcineurin acts via the C-terminus of NR2A to modulate desensitization of NMDA receptors. Neuropharmacology 42: 593-602.
  126. 126. Rycroft BK, Gibb AJ (2004) Inhibitory interactions of calcineurin (phosphatase 2B) and calmodulin on rat hippocampal NMDA receptors. Neuropharmacology 47: 505-514.
  127. 127. Lieberman DN, Mody I (1994) Regulation of NMDA channel function by endogenous Ca(2+)-dependent phosphatase. Nature 369: 235-239.
  128. 128. Hanney M, Prasher V, Williams N, Jones EL, Aarsland D, et al. (2012) Memantine for dementia in adults older than 40 years with Down's syndrome (MEADOWS): a randomised, double-blind, placebo-controlled trial. Lancet 379: 528-536.
  129. 129. Kleschevnikov AM, Belichenko PV, Gall J, George L, Nosheny R, et al. (2012) Increased efficiency of the GABAA and GABAB receptor-mediated neurotransmission in the Ts65Dn mouse model of Down syndrome. Neurobiol Dis 45: 683-691.
  130. 130. Best TK, Siarey RJ, Galdzicki Z (2007) Ts65Dn, a mouse model of Down syndrome, exhibits increased GABAB-induced potassium current. J Neurophysiol 97: 892-900.
  131. 131. Rissman RA, Mobley WC (2011) Implications for treatment: GABAA receptors in aging, Down syndrome and Alzheimer's disease. J Neurochem 117: 613-622.
  132. 132. Fernandez F, Morishita W, Zuniga E, Nguyen J, Blank M, et al. (2007) Pharmacotherapy for cognitive impairment in a mouse model of Down syndrome. Nat Neurosci 10: 411-413.
  133. 133. Rueda N, Florez J, Martinez-Cue C (2008) Chronic pentylenetetrazole but not donepezil treatment rescues spatial cognition in Ts65Dn mice, a model for Down syndrome. Neurosci Lett 433: 22-27.
  134. 134. Braudeau J, Dauphinot L, Duchon A, Loistron A, Dodd RH, et al. (2011) Chronic Treatment with a Promnesiant GABA-A alpha5-Selective Inverse Agonist Increases Immediate Early Genes Expression during Memory Processing in Mice and Rectifies Their Expression Levels in a Down Syndrome Mouse Model. Adv Pharmacol Sci 2011: 153218.
  135. 135. Clark S, Schwalbe J, Stasko MR, Yarowsky PJ, Costa AC (2006) Fluoxetine rescues deficient neurogenesis in hippocampus of the Ts65Dn mouse model for Down syndrome. Exp Neurol 200: 256-261.
  136. 136. Rueda N, Mostany R, Pazos A, Florez J, Martinez-Cue C (2005) Cell proliferation is reduced in the dentate gyrus of aged but not young Ts65Dn mice, a model of Down syndrome. Neurosci Lett 380: 197-201.
  137. 137. Ishihara K, Amano K, Takaki E, Shimohata A, Sago H, et al. (2010) Enlarged brain ventricles and impaired neurogenesis in the Ts1Cje and Ts2Cje mouse models of Down syndrome. Cereb Cortex 20: 1131-1143.
  138. 138. Whitaker-Azmitia PM (2001) Serotonin and brain development: role in human developmental diseases. Brain Res Bull 56: 479-485.
  139. 139. Bar-Peled O, Gross-Isseroff R, Ben-Hur H, Hoskins I, Groner Y, et al. (1991) Fetal human brain exhibits a prenatal peak in the density of serotonin 5-HT1A receptors. Neurosci Lett 127: 173-176.
  140. 140. Risser D, Lubec G, Cairns N, Herrera-Marschitz M (1997) Excitatory amino acids and monoamines in parahippocampal gyrus and frontal cortical pole of adults with Down syndrome. Life Sci 60: 1231-1237.
  141. 141. Banasr M, Hery M, Printemps R, Daszuta A (2004) Serotonin-induced increases in adult cell proliferation and neurogenesis are mediated through different and common 5-HT receptor subtypes in the dentate gyrus and the subventricular zone. Neuropsychopharmacology 29: 450-460.
  142. 142. Bianchi P, Ciani E, Guidi S, Trazzi S, Felice D, et al. (2010) Early pharmacotherapy restores neurogenesis and cognitive performance in the Ts65Dn mouse model for Down syndrome. J Neurosci 30: 8769-8779.
  143. 143. Boylan K, Romero S, Birmaher B (2007) Psychopharmacologic treatment of pediatric major depressive disorder. Psychopharmacology (Berl) 191: 27-38.
  144. 144. Bairy KL, Madhyastha S, Ashok KP, Bairy I, Malini S (2007) Developmental and behavioral consequences of prenatal fluoxetine. Pharmacology 79: 1-11.
  145. 145. Tyrrell J, Cosgrave M, McCarron M, McPherson J, Calvert J, et al. (2001) Dementia in people with Down's syndrome. Int J Geriatr Psychiatry 16: 1168-1174.
  146. 146. Millan Sanchez M, Heyn SN, Das D, Moghadam S, Martin KJ, et al. (2011) Neurobiological elements of cognitive dysfunction in down syndrome: exploring the role of APP. Biol Psychiatry 71: 403-409.
  147. 147. Rovelet-Lecrux A, Hannequin D, Raux G, Le Meur N, Laquerriere A, et al. (2006) APP locus duplication causes autosomal dominant early-onset Alzheimer disease with cerebral amyloid angiopathy. Nat Genet 38: 24-26.
  148. 148. Prasher VP, Farrer MJ, Kessling AM, Fisher EM, West RJ, et al. (1998) Molecular mapping of Alzheimer-type dementia in Down's syndrome. Ann Neurol 43: 380-383.
  149. 149. Fischer DF, van Dijk R, Sluijs JA, Nair SM, Racchi M, et al. (2005) Activation of the Notch pathway in Down syndrome: cross-talk of Notch and APP. Faseb J 19: 1451-1458.
  150. 150. Stockley JH, O'Neill C (2007) The proteins BACE1 and BACE2 and beta-secretase activity in normal and Alzheimer's disease brain. Biochem Soc Trans 35: 574-576.
  151. 151. Ahmed RR, Holler CJ, Webb RL, Li F, Beckett TL, et al. (2010) BACE1 and BACE2 enzymatic activities in Alzheimer's disease. J Neurochem 112: 1045-1053.
  152. 152. Holler CJ, Webb RL, Laux AL, Beckett TL, Niedowicz DM, et al. (2012) BACE2 expression increases in human neurodegenerative disease. Am J Pathol 180: 337-350.
  153. 153. Ryoo SR, Cho HJ, Lee HW, Jeong HK, Radnaabazar C, et al. (2008) Dual-specificity tyrosine(Y)-phosphorylation regulated kinase 1A-mediated phosphorylation of amyloid precursor protein: evidence for a functional link between Down syndrome and Alzheimer's disease. J Neurochem 104: 1333-1344.
  154. 154. Lloret A, Badia MC, Giraldo E, Ermak G, Alonso MD, et al. (2011) Amyloid-beta toxicity and tau hyperphosphorylation are linked via RCAN1 in Alzheimer's disease. J Alzheimers Dis 27: 701-709.
  155. 155. Jung MS, Park JH, Ryu YS, Choi SH, Yoon SH, et al. (2011) Regulation of RCAN1 protein activity by Dyrk1A protein-mediated phosphorylation. J Biol Chem 286: 40401-40412.
  156. 156. Hirotsune S, Yoshida N, Chen A, Garrett L, Sugiyama F, et al. (2003) An expressed pseudogene regulates the messenger-RNA stability of its homologous coding gene. Nature 423: 91-96.
  157. 157. Magistri M, Faghihi MA, St Laurent G, 3rd, Wahlestedt C (2012) Regulation of chromatin structure by long noncoding RNAs: focus on natural antisense transcripts. Trends Genet 28: 389-396.
  158. 158. Kurokawa R (2011) Promoter-associated long noncoding RNAs repress transcription through a RNA binding protein TLS. Adv Exp Med Biol 722: 196-208.
  159. 159. Kogo R, Shimamura T, Mimori K, Kawahara K, Imoto S, et al. (2011) Long noncoding RNA HOTAIR regulates polycomb-dependent chromatin modification and is associated with poor prognosis in colorectal cancers. Cancer Res 71: 6320-6326.
  160. 160. Yoon JH, Abdelmohsen K, Srikantan S, Yang X, Martindale JL, et al. (2012) LincRNA-p21 Suppresses Target mRNA Translation. Mol Cell.
  161. 161. Huarte M, Guttman M, Feldser D, Garber M, Koziol MJ, et al. (2010) A large intergenic noncoding RNA induced by p53 mediates global gene repression in the p53 response. Cell 142: 409-419.
  162. 162. Kuhn DE, Nuovo GJ, Terry AV, Jr., Martin MM, Malana GE, et al. (2010) Chromosome 21-derived microRNAs provide an etiological basis for aberrant protein expression in human Down syndrome brains. J Biol Chem 285: 1529-1543.
  163. 163. Robinson L, Guy J, McKay L, Brockett E, Spike RC, et al. (2012) Morphological and functional reversal of phenotypes in a mouse model of Rett syndrome. Brain.
  164. 164. Weng SM, Bailey ME, Cobb SR (2011) Rett syndrome: from bed to bench. Pediatr Neonatol 52: 309-316.
  165. 165. Malatkova P, Maser E, Wsol V (2010) Human carbonyl reductases. Curr Drug Metab 11: 639-658.
  166. 166. Rachidi M, Delezoide AL, Delabar JM, Lopes C (2009) A quantitative assessment of gene expression (QAGE) reveals differential overexpression of DOPEY2, a candidate gene for mental retardation, in Down syndrome brain regions. Int J Dev Neurosci 27: 393-398.
  167. 167. Mimura Y, Takahashi K, Kawata K, Akazawa T, Inoue N (2010) Two-step colocalization of MORC3 with PML nuclear bodies. J Cell Sci 123: 2014-2024.
  168. 168. Mello JA, Sillje HH, Roche DM, Kirschner DB, Nigg EA, et al. (2002) Human Asf1 and CAF-1 interact and synergize in a repair-coupled nucleosome assembly pathway. EMBO Rep 3: 329-334.
  169. 169. Wilcox ER, Burton QL, Naz S, Riazuddin S, Smith TN, et al. (2001) Mutations in the gene encoding tight junction claudin-14 cause autosomal recessive deafness DFNB29. Cell 104: 165-172.
  170. 170. Bao B, Pestinger V, Hassan YI, Borgstahl GE, Kolar C, et al. (2011) Holocarboxylase synthetase is a chromatin protein and interacts directly with histone H3 to mediate biotinylation of K9 and K18. J Nutr Biochem 22: 470-475.
  171. 171. Shibuya K, Kudoh J, Minoshima S, Kawasaki K, Asakawa S, et al. (2000) Isolation of two novel genes, DSCR5 and DSCR6, from Down syndrome critical region on human chromosome 21q22.2. Biochem Biophys Res Commun 271: 693-698.
  172. 172. Watanabe R, Murakami Y, Marmor MD, Inoue N, Maeda Y, et al. (2000) Initial enzyme for glycosylphosphatidylinositol biosynthesis requires PIG-P and is regulated by DPM2. Embo J 19: 4402-4411.
  173. 173. Nakamura A, Hattori M, Sakaki Y (1997) Isolation of a novel human gene from the Down syndrome critical region of chromosome 21q22.2. J Biochem 122: 872-877.
  174. 174. Du Y, Zhang J, Wang H, Yan X, Yang Y, et al. (2011) Hypomethylated DSCR4 is a placenta-derived epigenetic marker for trisomy 21. Prenat Diagn 31: 207-214.
  175. 175. de Wit NJ, Cornelissen IM, Diepstra JH, Weidle UH, Ruiter DJ, et al. (2005) The MMA1 gene family of cancer-testis antigens has multiple alternative splice variants: characterization of their expression profile, the genomic organization, and transcript properties. Genes Chromosomes Cancer 42: 10-21.
  176. 176. Okamoto K, Iwasaki N, Doi K, Noiri E, Iwamoto Y, et al. (2012) Inhibition of Glucose-Stimulated Insulin Secretion by KCNJ15, a Newly Identified Susceptibility Gene for Type 2 Diabetes. Diabetes 61: 1734-1741.
  177. 177. Hirano Y, Hendil KB, Yashiroda H, Iemura S, Nagane R, et al. (2005) A heterodimeric complex that promotes the assembly of mammalian 20S proteasomes. Nature 437: 1381-1385.
  178. 178. Vilardi F, Lorenz H, Dobberstein B (2011) WRB is the receptor for TRC40/Asna1-mediated insertion of tail-anchored proteins into the ER membrane. J Cell Sci 124: 1301-1307.
  179. 179. den Hollander AI, Koenekoop RK, Mohamed MD, Arts HH, Boldt K, et al. (2007) Mutations in LCA5, encoding the ciliary protein lebercilin, cause Leber congenital amaurosis. Nat Genet 39: 889-895.
  180. 180. Seko A, Kataoka F, Aoki D, Sakamoto M, Nakamura T, et al. (2009) Beta1,3-galactosyltransferases-4/5 are novel tumor markers for gynecological cancers. Tumour Biol 30: 43-50.
  181. 181. Hirabayashi S, Tajima M, Yao I, Nishimura W, Mori H, et al. (2003) JAM4, a junctional cell adhesion molecule interacting with a tight junction protein, MAGI-1. Mol Cell Biol 23: 4267-4282.
  182. 182. Harashima S, Wang Y, Horiuchi T, Seino Y, Inagaki N (2011) Purkinje cell protein 4 positively regulates neurite outgrowth and neurotransmitter release. J Neurosci Res 89: 1519-1530.
  183. 183. Xu Y, Ye H, Shen Y, Xu Q, Zhu L, et al. (2011) Dscam mutation leads to hydrocephalus and decreased motor function. Protein Cell 2: 647-655.
  184. 184. Robert-Cooperman CE, Carnegie JR, Wilson CG, Yang J, Cook JR, et al. (2010) Targeted disruption of pancreatic-derived factor (PANDER, FAM3B) impairs pancreatic beta-cell function. Diabetes 59: 2209-2218.
  185. 185. Muller M, Winnacker EL, Brem G (1992) Molecular cloning of porcine Mx cDNAs: new members of a family of interferon-inducible proteins with homology to GTP-binding proteins. J Interferon Res 12: 119-129.
  186. 186. Kelso J, Visagie J, Theiler G, Christoffels A, Bardien S, et al. (2003) eVOC: a controlled vocabulary for unifying gene expression data. Genome Res 13: 1222-1230.

Written By

Ferdinando Di Cunto and Gaia Berto

Submitted: 01 May 2012 Published: 06 March 2013