Skip to main content
Cellular and Molecular Life Sciences: CMLS logoLink to Cellular and Molecular Life Sciences: CMLS
. 2009 Aug 14;66(20):3235–3240. doi: 10.1007/s00018-009-0123-2

Function and regulation of Dyrk1A: towards understanding Down syndrome

Joongkyu Park 1, Woo-Joo Song 2,, Kwang Chul Chung 1,
PMCID: PMC11115655  PMID: 19685005

Abstract

Down syndrome (DS) is associated with a variety of symptoms, such as incapacitating mental retardation and neurodegeneration (i.e., Alzheimer’s disease), that prevent patients from leading fully independent lives. These phenotypes are a direct consequence of the overexpression of chromosome 21 genes, which are present in duplicate due to non-disjunction of chromosome 21. Accumulating data suggest that the chromosome 21 gene product, dual-specificity tyrosine-(Y)-phosphorylation regulated kinase 1A (Dyrk1A), participates in the pathogenic mechanisms underlying the mental and other physical symptoms of DS. In this review, we summarize the evidence supporting a role for Dyrk1A in DS, especially DS pathogenesis. Recently, several natural and synthetic compounds have been identified as Dyrk1A inhibitors. Understanding the function and regulation of Dyrk1A may lead to the development of novel therapeutic agents aimed at treating DS.

Keywords: Down syndrome, Dyrk1A, Kinase, Phosphorylation, Inhibitor

Introduction

Down syndrome (DS) is one of the most common genetic defects, with an incidence of 1 in every ~700 births. This condition arises from a complete or partial duplication of human chromosome 21 (trisomy 21) [1]. The resulting imbalance in gene expression causes more than 20 neural and non-neural symptoms, including a distinct facial appearance, hypotonia, congenital heart defects, mental retardation, early onset Alzheimer’s disease (AD), susceptibility to leukemia, gastrointestinal malformations, and immune system defects [2]. To date, more than 530 genes (176 conserved, 355 non-conserved) have been identified on chromosome 21, making identification of the genes responsible for specific DS phenotypes a daunting task [3]. Studies of rare cases of partial trisomy 21 suggest that a defined chromosome 21 region called the Down syndrome critical region (DSCR, 21q22.1-22.3) may cause the typical features of DS [46]. Among more than 30 presumed genes located on the DSCR, the dual-specificity tyrosine-(Y)-regulated kinase 1A gene (Dyrk1A) is associated with some DS characteristics, mental retardation, and motor defects [79], as well as with neurodegenerative diseases such as AD [1012], Parkinson’s disease (PD), and Huntington’s disease (HD) [1315].

Dyrk1A kinase in DS: regulation and function

Drosophila melanogaster minibrain protein kinase, a homolog of Dyrk1A, is essential for normal postembryonic neurogenesis [16]. Mutant flies expressing lower levels of minibrain exhibit a marked size reduction in the optic lobes and central brain, as well as visual, olfactory and motor defects. The human orthologue may have similar functions that are important to DS phenotypes. This has prompted several research laboratories to isolate mammalian counterparts (i.e., human, rat, and mouse Dyrk1A) that are highly conserved (>99% identity) across all 763 residues [1719]. The Dyrk family consists of five mammalian members (Dyrk1A, Dyrk1B, Dyrk2, Dyrk3, and Dyrk4). Among them, Dyrk1A is the only member located on chromosome 21 [20]. Dyrk1A is a proline-directed protein kinase that contains multiple domains, including a nuclear localization signal at the N-terminus, a kinase domain, a PEST domain for protein degradation, a 13-consecutive-histidine repeat, and an S/T-rich region that has an unknown function (Fig. 1). Outside the kinase domain, Dyrk1 does not share significant sequence homology with other family members.

Fig. 1.

Fig. 1

Domain structure of Dyrk1A and other Dyrk family members. NLS Nuclear localization signal, KINASE kinase domain, PEST (Pro, Glu, Ser, Thr)-rich domain, His 13-consecutive-histidine repeat, S/T (Ser, Thr)-rich region

As its name implies, Dyrk1A has dual substrate specificities. Dyrk1A undergoes self-activation through autophosphorylation at Tyr 321, which is situated in the kinase domain [18]. Autophosphorylation at Tyr 321 is thought to occur during protein synthesis and through the intramolecular formation of a transitory intermediate, producing a constitutively active form of Dyrk1A [21]. Dyrk1A also phosphorylates target substrates at Ser or Thr residue. Dyrk1A has been shown to phosphorylate or interact with more than two dozen proteins, which are associated with multiple pathways. Even though Dyrk1A has a nuclear localization signal and 13-histidine repeat for nuclear speckle targeting [18, 22], it has also been detected in the soma and dendrites of neurons [23]. Therefore, it is not surprising that Dyrk1A substrates comprise both nuclear and cytosolic proteins, including transcriptional factors (CREB, NFAT, STAT3, FKHR, Gli1), splicing factors (cyclin L2, SF2, SF3), a translation factor (eIF2Bε), synaptic proteins (dynamin I, amphiphysin I, synaptojanin I), and miscellaneous proteins (glycogen synthase, caspase-9, Notch). This substrate diversity points to pleiotropic roles for Dyrk1A [2428] (Fig. 2). In particular, the involvement of Dyrk1A in the regulation of NFAT pathway plays a role in the cell cycle control and synaptic function [29, 30], and the 1.5-fold increase of Dyrk1A gene in mice reduced NFAT transcriptional activity and caused dysregulated vertebrate development, including vascular defects and failed heart valve development [29].

Fig. 2.

Fig. 2

Substrates and multiple putative roles of Dyrk1A in gene transcription, mRNA splicing, synapse function, and neurodegeneration

Alterations in Dyrk1A expression are frequently associated with DS phenotypes. Dyrk1A mRNA is overexpressed in DS fetal brains and Ts65Dn mice, a well-established murine model for DS [31]. Dyrk1A is elevated approximately 1.5-fold in a gene dosage-dependent manner in DS patients [32]. Clues as to the cellular function of Dyrk1A may be provided by Dyrk1A expression patterns under normal conditions. Dyrk1A is expressed ubiquitously, although not evenly, in fetal and adult tissues as well as in non-neuronal tissues and the central nervous system. Particularly strong expression has been noted in the cerebellum, olfactory bulb, and hippocampus [17, 19]. In the mouse brain, Dyrk1A is present in preneurogenic progenitors as early as embryonic day (E) 8.5, indicating that Dyrk1A participates in early development [23]. The strong Dyrk1A expression seen in mice at embryonic stages and at birth gradually decreases until lower expression levels stabilize at around 3 weeks (Song, unpublished observation). Since over- or underexpression of constitutively active Dyrk1A in transgenic mice produces phenotypes seen in DS, the cellular function of Dyrk1A is presumed to be sensitive to its expression and/or degradation. Dyrk1A expression can be enhanced by treatment with β-amyloid (Aβ) or overexpression of the transcription factor E2F1 [33, 34]. On the other hand, it can be repressed by the AP4-geminin complex [35]. In addition, Dyrk1A activity can be increased by binding of 14-3-3 [36].

Animal models with altered Dyrk1A expression

Several murine models for DS with segmental trisomy of mouse chromosome 16 (MMU16), which corresponds to human chromosome 21, have been developed by chromosome engineering. There are three representative DS model mice bearing the Dyrk1A gene: Ts65Dn, Ts1Cje, and Ts1Rhr. Ts65Dn mice contain an extra segment of a distal ~17 Mb region (~104 genes) of MMU16 [37], while Ts1Cje mice have an extra distal ~8.3 Mb region (~81 genes) [38]. These two models display a number of developmental and neuropathological characteristics similar to DS patients, including learning and behavior abnormalities [37, 38], altered synaptic plasticity [39, 40], and changes in dendritic spines of hippocampus and cortex [41, 42]. Recently, Ts1Rhr mice, which contain an even smaller trisomic segment of MMU16 (~33 genes), were reported to exhibit cognitive and behavior abnormalities and changes in spine density and morphology [43], although they seemed to have normal hippocampal function [44]. Thus, the specific genes on DSCR are suspected of DS pathogenesis. Moreover, the gene dosage effect in DS was further supported by the finding that the pathologic features of DS model mice were restored back to normal when the critical region genes are returned to their normal dosage [44].

Pertaining to Dyrk1A, its knock-out mice are embryonic lethal (E14.5), and heterozygous Dyrk1A mice (Dyrk1A +/−) exhibit decreased neonatal viability, developmental delay, and altered neocortical pyramidal cells. Thus, Dyrk1A has vital and sensitive gene dosage effects [45, 46]. Also consistent with a role for Dyrk1A in neurodevelopment, microcephaly has been reported in two unrelated patients with Dyrk1A truncation [47]. Mouse models overexpressing Dyrk1A have also been produced. Mice carry the human Dyrk1A genomic DNA in a bacterial or yeast artificial chromosome, or they carry extra copies of murine Dyrk1A cDNA [79]. These transgenic animals exhibit hippocampal-dependent spatial learning and memory deficit in the Morris water maze, developmental delay, and motor deficits, strongly implicating Dyrk1A overexpression in several DS phenotypes.

Role of Dyrk1A in the pathogenesis of other diseases

In addition to participating in DS, Dyrk1A appears to be involved in the pathogenesis of several neurodegenerative diseases such as AD, PD, and HD (Fig. 2). Owing to improvements in living conditions and advances in medical care, DS patients are expected to live up to the end of their sixth decade and in some cases, beyond this age. However, adults with DS typically undergo premature aging and exhibit a rapid decline in memory ability. They are also particularly vulnerable to AD [1012]. Nearly all DS patients develop AD-like dementia several decades earlier than does the general population. By the age of 40, these patients have the neuropathological lesions seen in AD as well as dementia [48, 49]. The pathogenic DS brains have AD hallmarks including amyloid plaques and neurofibrillary tangles (i.e., insoluble deposits consisting of Aβ and abnormally hyperphosphorylated Tau, respectively). Dyrk1A phosphorylates Thr 668 of amyloid precursor protein. This may increase Aβ production, which in turn enhances Dyrk1A expression [34, 50, 51]. Dyrk1A can also phosphorylate several critical Tau residues in tangles, and it is present in tangles and the sarkosyl insoluble fraction of AD brains [50, 5254]. Taken together, these findings suggest that the overexpression of Dyrk1A in DS may play a role in accelerating AD pathogenesis. Additional evidence suggests that Dyrk1A participates in not only AD, but also PD. Dyrk1A phosphorylation of α-synuclein, a key Lewy body component, enhances the α-synuclein positive inclusion that leads to neuronal cell death [14]. Involvement of Dyrk1A in AD and PD has been further corroborated by the finding that the GTPase septin 4 present in AD tangles and PD inclusions can be phosphorylated by Dyrk1A [15]. We have previously found that Dyrk1A phosphorylation of huntingtin-interacting protein-1 modulates differentiation and death in hippocampal neuroprogenitor cells, suggesting that Dyrk1A participates in HD pathogenesis [13]. In addition, Dyrk1A affects neuronal proliferation by phosphorylating p53 during embryonic brain development (Park and Chung, submitted).

Therapeutic approaches

Mental retardation affects 1–2% of the general population, and a large percentage of affected individuals suffer from chromosomal abnormalities. Mental retardation caused by DS contributes to about 30% of these cases, making DS the most common cause of mental retardation [55, 56]. For DS patients, mental retardation is the greatest obstacle preventing normal, independent living. Therapeutic agents aimed at treating DS phenotypes by targeting the responsible protein(s) have yet to be developed due to the lack of characterization of the responsible gene(s). Several AD drugs (e.g., acetylcholine esterase inhibitors) have been used to enhance the cognitive function of adult DS patients since DS has neuropathological similarities with AD [57]. However, the results are controversial, ranging from a significant improvement in dementia scores to adverse effects or no improvement at all [58, 59]. A similar controversy exists with regard to efficacy and the treatment of DS children with Piracetam, a nootropic agent known to enhance cognitive performance [60]. Recent attempts to ameliorate the abnormalities of learning and behavior in Ts65Dn mice revealed that non-competitive or competitive antagonists of GABAA receptor (picrotoxin, bilobalide, pentylenetetrazole) and NMDA receptor (MK-801, memantine) can rescue them from the defects [61, 62].

Recently, a number of researchers have investigated whether Dyrk1A-associated learning and memory deficits can be alleviated by modulating Dyrk1A expression or activity. Several approaches such as the use of natural products, synthetic inhibitors, or shRNA have been investigated. For example, treatment of Dyrk1A transgenic mice with epigallocatechin-3-gallate, a major polyphenolic constituent of green tea, rescues alterations in brain volume and ameliorates cognitive deficits [63]. In addition, harmine, which is found in the Middle Eastern plant (i.e., Peganum hamarla) and the South American vine (i.e., Banisteriopsis caapi), inhibits the kinase activity of Dyrk1A at nanomolar range [64]. Injection of adeno-associated virus vector that encodes inhibitory Dyrk1A shRNA into the striata of Dyrk1A transgenic mice restores motor coordination, attenuates hyperactivity, and improves sensorimotor gating [65]. Several synthetic Dyrk1A inhibitors have also been isolated, although their in vivo efficacy has yet to be tested [66, 67].

Closing remarks

Overexpression of genes present on an additional copy of chromosome 21 causes multiple phenotypes seen in DS. Among DS symptoms, mental retardation and early onset AD hinder the normal daily lives of DS patients during early and middle age. The chromosome 21 gene product, Dyrk1A protein kinase, is overexpressed in DS patients and has recently attracted considerable attention due to its association with several DS phenotypes, including mental retardation and AD. The finding that Dyrk1A has more than two dozen substrates in multiple pathways suggests that Dyrk1A participates in multiple DS symptoms as well as neurodegenerative PD and HD. Studies by our laboratory and others show that Dyrk1A functions as a negative regulator of cardiomyocyte hypertrophy [68], and that it participates in bone homeostasis (Lee et al., manuscript in revision) and abnormal immune responses via NFAT phosphorylation (Song et al., unpublished observation). This suggests that Dyrk1A overexpression is responsible for other DS phenotypes such as congenital heart defects, short stature, and immune system defects.

Nonetheless, multiple genes could be involved in the genesis of DS, and there are several examples indicating other genes alone or together in the DSCR could contribute to DS pathogenesis. For example, the 1.5-fold increase of the DSCR1 gene caused the similar consequences to that observed in Dyrk1A transgenic mice by inhibiting calcineurin and eventually increasing NFAT phosphorylation [29]. DSCR1 transgenic mice also displayed suppression of tumor growth, one of the typical features seen in DS patients, and this suppression was further enhanced by combinatorial overexpression of Dyrk1A [69]. Although trials have been conducted to test the efficacy of AD drugs in treating DS symptoms, novel therapeutic agents are not yet available to DS patients. The development of therapeutic agents that effectively repress Dyrk1A expression or activity is in its infancy. Nevertheless, the results of initial attempts to treat Dyrk1A transgenic mice using this approach are encouraging and suggest that it holds promise for the treatment of DS phenotypes and other neurodegenerative diseases. Knowledge of the mechanisms triggered by Dyrk1A overexpression will bring us closer to understanding DS and developing effective therapies for this condition.

Acknowledgments

This work was supported by the Korea Science and Engineering Foundation (KOSEF) grants (R01-2007-000-11910-0 to W.-J.S. and R11-2007-040-01005-0 to K.C.C.) funded by the Ministry of Education, Science and Technology (MEST), a Korea Research Foundation Grant funded by the Korean Government (MOEHRD) (KRF-2008-314-E00180 to W.-J.S.), and the Korea Health 21 R&D Project (A080551 to K.C.C.) funded by the Ministry of Health, Welfare and Family Affairs. This work was also partly supported by KOSEF grant through the National Research Laboratory Program (R04-2007-000-20014-0 to K.C.C.).

Contributor Information

Woo-Joo Song, Phone: +82-42-8924186, FAX: +82-51-8920059, Email: wjsong@inje.ac.kr.

Kwang Chul Chung, Phone: +82-2-21232653, FAX: +82-2-3125657, Email: kchung@yonsei.ac.kr.

References

  • 1.Lejeune J, Gautier M, Turpin R. Study of somatic chromosomes from 9 mongoloid children. C R Hebd Seances Acad Sci. 1959;248:1721–1722. [PubMed] [Google Scholar]
  • 2.Korenberg JR, Chen X-N, Schipper R, Sun Z, Gonsky R, Gerwehr S, Carpenter N, Daumer C, Dignan P, Disteche C, Graham JM, Jr, Hugdins L, McGillivray B, Miyazaki K, Ogasawara N, Park JP, Pagon R, Pueschel S, Sack G, Say B, Schuffenhauer S, Soukup S, Yamanaka T. Down syndrome phenotypes: the consequences of chromosomal imbalance. Proc Natl Acad Sci USA. 1994;91:4997–5001. doi: 10.1073/pnas.91.11.4997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Gardiner K (2007) Overview of the genes of the human chromosome 21. In: Pritchard M, Reeves RH, Dierssenc M, Patterson D, Gardiner KJ (eds) Down syndrome and the genes of human chromosome 21: current knowledge and future potentials. Report on the expert workshop on the biology of chromosome 21 genes: towards gene-phenotype correlations in Down syndrome. September 28-October 1, 2007, Washington D.C. [DOI] [PubMed]
  • 4.Korenberg JR, Kawashima H, Pulst S-M, Ikeuchi T, Ogasawara N, Yamamoto K, Schonberg SA, West R, Allen L, Magenis E, Ikawa K, Taniguchi N, Epstein CJ. Molecular definition of a region of chromosome 21 that causes features of the Down syndrome phenotype. Am J Hum Genet. 1990;47:236–246. [PMC free article] [PubMed] [Google Scholar]
  • 5.McCormick MK, Schinzel A, Petersen MB, Stetten G, Driscoll DJ, Cantu ES, Tranebjaerg L, Mikkelsen M, Watkins PC, Antonarakis SE. Molecular genetic approach to the characterization of the “Down syndrome region” of chromosome 21. Genomics. 1989;5:325–331. doi: 10.1016/0888-7543(89)90065-7. [DOI] [PubMed] [Google Scholar]
  • 6.Rahmani Z, Blouin J-L, Creau-Goldberg N, Watkins PC, Mattei J-F, Poissonnier M, Prieur M, Chettouh Z, Nicole A, Aurias A, Sinet P-M, Delabar J-M. Critical role of the D21S55 region on chromosome 21 in the pathogenesis of Down syndrome. Proc Natl Acad Sci USA. 1989;86:5958–5962. doi: 10.1073/pnas.86.15.5958. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Ahn K-J, Jeong HK, Choi H-S, Ryoo S-R, Kim YJ, Goo J-S, Choi S-Y, Han J-S, Ha I, Song W-J. DYRK1A BAC transgenic mice show altered synaptic plasticity with learning and memory defects. Neurobiol Dis. 2006;22:463–472. doi: 10.1016/j.nbd.2005.12.006. [DOI] [PubMed] [Google Scholar]
  • 8.Altafaj X, Dierssen M, Baamonde C, Marti E, Visa J, Guimera J, Oset M, Gonzalez JR, Florez J, Fillat C, Estivill X. Neurodevelopmental delay, motor abnormalities and cognitive deficits in transgenic mice overexpressing Dyrk1A (minibrain), a murine model of Down’s syndrome. Hum Mol Genet. 2001;10:1915–1923. doi: 10.1093/hmg/10.18.1915. [DOI] [PubMed] [Google Scholar]
  • 9.Smith DJ, Stevens ME, Sudanagunta SP, Bronson RT, Makhinson M, Watabe AM, O’Dell TJ, Fung J, Weier HU, Cheng JF, Rubin EM. Functional screening of 2 Mb of human chromosome 21q22.2 in transgenic mice implicates minibrain in learning defects associated with Down syndrome. Nat Genet. 1997;16:28–36. doi: 10.1038/ng0597-28. [DOI] [PubMed] [Google Scholar]
  • 10.Burger PC, Vogel FS. The development of the pathologic changes of Alzheimer’s disease and senile dementia in patients with Down’s syndrome. Am J Pathol. 1973;73:457–476. [PMC free article] [PubMed] [Google Scholar]
  • 11.Olson MI, Shaw CM. Presenile dementia and Alzheimer’s disease in mongolism. Brain. 1969;92:147–156. doi: 10.1093/brain/92.1.147. [DOI] [PubMed] [Google Scholar]
  • 12.Wisniewski KE, Wisniewski HM, Wen GY. Occurrence of neuropathological changes and dementia of Alzheimer’s disease in Down’s syndrome. Ann Neurol. 1985;17:278–282. doi: 10.1002/ana.410170310. [DOI] [PubMed] [Google Scholar]
  • 13.Kang JE, Choi SA, Park JB, Chung KC. Regulation of the proapoptotic activity of huntingtin interacting protein 1 by Dyrk1 and caspase-3 in hippocampal neuroprogenitor cells. J Neurosci Res. 2005;81:62–72. doi: 10.1002/jnr.20534. [DOI] [PubMed] [Google Scholar]
  • 14.Kim EJ, Sung JY, Lee HJ, Rhim H, Hasegawa M, Iwatsubo T, Min DS, Kim J, Paik SR, Chung KC. Dyrk1A phosphorylates alpha-synuclein and enhances intracellular inclusion formation. J Biol Chem. 2006;281:33250–33257. doi: 10.1074/jbc.M606147200. [DOI] [PubMed] [Google Scholar]
  • 15.Sitz JH, Baumgartel K, Hammerle B, Papadopoulos C, Hekerman P, Tejedor FJ, Becker W, Lutz B. The Down syndrome candidate dual-specificity tyrosine phosphorylation-regulated kinase 1A phosphorylates the neurodegeneration-related septin 4. Neuroscience. 2008;157:596–605. doi: 10.1016/j.neuroscience.2008.09.034. [DOI] [PubMed] [Google Scholar]
  • 16.Tejedor F, Zhu XR, Kaltenbach E, Ackermann A, Baumann A, Canal I, Heisenberg M, Fischbach KF, Pongs O. Minibrain: a new protein kinase family involved in postembryonic neurogenesis in Drosophila . Neuron. 1995;14:287–301. doi: 10.1016/0896-6273(95)90286-4. [DOI] [PubMed] [Google Scholar]
  • 17.Guimera J, Casas C, Pucharcos C, Solans A, Domenech A, Planas AM, Ashley J, Lovett M, Estivill X, Pritchard MA. A human homologue of Drosophila minibrain (MNB) is expressed in the neuronal regions affected in Down syndrome and maps to the critical region. Hum Mol Genet. 1996;5:1305–1310. doi: 10.1093/hmg/5.9.1305. [DOI] [PubMed] [Google Scholar]
  • 18.Kentrup H, Becker W, Heukelbach J, Wilmes A, Schurmann A, Huppertz C, Kainulainen H, Joost HG. Dyrk, a dual specificity protein kinase with unique structural features whose activity is dependent on tyrosine residues between subdomains VII and VIII. J Biol Chem. 1996;271:3488–3495. doi: 10.1074/jbc.271.7.3488. [DOI] [PubMed] [Google Scholar]
  • 19.Song W-J, Sternberg LR, Kasten-Sportes C, Keuren ML, Chung S-H, Slack AC, Miller DE, Glover TW, Chiang P-W, Lou L, Kurnit DM. Isolation of human and murine homologues of the Drosophila minibrain gene: human homologue maps to 21q22.2 in the Down syndrome “critical region”. Genomics. 1996;38:331–339. doi: 10.1006/geno.1996.0636. [DOI] [PubMed] [Google Scholar]
  • 20.Becker W, Joost HG. Structural and functional characteristics of Dyrk, a novel subfamily of protein kinases with dual specificity. Prog Nucleic Acid Res Mol Biol. 1999;62:1–17. doi: 10.1016/S0079-6603(08)60503-6. [DOI] [PubMed] [Google Scholar]
  • 21.Lochhead PA, Sibbet G, Morrice N, Cleghon V. Activation-loop autophosphorylation is mediated by a novel transitional intermediate form of DYRKs. Cell. 2005;121:925–936. doi: 10.1016/j.cell.2005.03.034. [DOI] [PubMed] [Google Scholar]
  • 22.Alvarez M, Estivill X, de la Luna S. DYRK1A accumulates in splicing speckles through a novel targeting signal and induces speckle disassembly. J Cell Sci. 2003;116:3099–3107. doi: 10.1242/jcs.00618. [DOI] [PubMed] [Google Scholar]
  • 23.Hammerle B, Elizalde C, Tejedor FJ. 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. 2008;27:1061–1074. doi: 10.1111/j.1460-9568.2008.06092.x. [DOI] [PubMed] [Google Scholar]
  • 24.Galceran J, de Graaf K, Tejedor FJ, Becker W. The MNB/DYRK1A protein kinase: genetic and biochemical properties. J Neural Transm Suppl. 2003;67:139–148. doi: 10.1007/978-3-7091-6721-2_12. [DOI] [PubMed] [Google Scholar]
  • 25.Hammerle B, Elizalde C, Galceran J, Becker W, Tejedor FJ. The MNB/DYRK1A protein kinase: neurobiological functions and Down syndrome implications. J Neural Transm Suppl. 2003;67:129–137. doi: 10.1007/978-3-7091-6721-2_11. [DOI] [PubMed] [Google Scholar]
  • 26.Skurat AV, Dietrich AD. Phosphorylation of Ser640 in muscle glycogen synthase by DYRK family protein kinases. J Biol Chem. 2004;279:2490–2498. doi: 10.1074/jbc.M301769200. [DOI] [PubMed] [Google Scholar]
  • 27.Laguna A, Aranda S, Barallobre MJ, Barhoum R, Fernandez E, Fotaki V, Delabar JM, de la Luna S, de la Villa P, Arbones ML. The protein kinase DYRK1A regulates caspase-9-mediated apoptosis during retina development. Dev Cell. 2008;15:841–853. doi: 10.1016/j.devcel.2008.10.014. [DOI] [PubMed] [Google Scholar]
  • 28.Fernandez-Martinez J, Vela EM, Tora-Ponsioen M, Ocana OH, Nieto MA, Galceran J. Attenuation of Notch signalling by the Down-syndrome-associated kinase DYRK1A. J Cell Sci. 2009;122:1574–1583. doi: 10.1242/jcs.044354. [DOI] [PubMed] [Google Scholar]
  • 29.Arron JR, Winslow MM, Polleri A, Chang CP, Wu H, Gao X, Neilson JR, Chen L, Heit JJ, Kim SK, Yamasaki N, Miyakawa T, Francke U, Graef IA, Crabtree GR. NFAT dysregulation by increased dosage of DSCR1 and DYRK1A on chromosome 21. Nature. 2006;441:595–600. doi: 10.1038/nature04678. [DOI] [PubMed] [Google Scholar]
  • 30.Gwack Y, Sharma S, Nardone J, Tanasa B, Iuga A, Srikanth S, Okamura H, Bolton D, Feske S, Hogan PG, Rao A. A genome-wide Drosophila RNAi screen identifies DYRK-family kinases as regulators of NFAT. Nature. 2006;441:646–650. doi: 10.1038/nature04631. [DOI] [PubMed] [Google Scholar]
  • 31.Guimera J, Casas C, Estivill X, Pritchard M. Human minibrain homologue (MNBH/DYRK1): characterization, alternative splicing, differential tissue expression, and overexpression in Down syndrome. Genomics. 1999;57:407–418. doi: 10.1006/geno.1999.5775. [DOI] [PubMed] [Google Scholar]
  • 32.Dowjat WK, Adayev T, Kuchna I, Nowicki K, Palminiello S, Hwang YW, Wegiel J. Trisomy-driven overexpression of DYRK1A kinase in the brain of subjects with Down syndrome. Neurosci Lett. 2007;413:77–81. doi: 10.1016/j.neulet.2006.11.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Kimura R, Kamino K, Yamamoto M, Nuripa A, Kida T, Kazui H, Hashimoto R, Tanaka T, Kudo T, Yamagata H, Tabara Y, Miki T, Akatsu H, Kosaka K, Funakoshi E, Nishitomi K, Sakaguchi G, Kato A, Hattori H, Uema T, Takeda M. The DYRK1A gene, encoded in chromosome 21 Down syndrome critical region, bridges between beta-amyloid production and tau phosphorylation in Alzheimer disease. Hum Mol Genet. 2007;16:15–23. doi: 10.1093/hmg/ddl437. [DOI] [PubMed] [Google Scholar]
  • 34.Maenz B, Hekerman P, Vela EM, Galceran J, Becker W. Characterization of the human DYRK1A promoter and its regulation by the transcription factor E2F1. BMC Mol Biol. 2008;9:30. doi: 10.1186/1471-2199-9-30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Kim MY, Jeong BC, Lee JH, Kee HJ, Kook H, Kim NS, Kim YH, Kim JK, Ahn KY, Kim KK. A repressor complex, AP4 transcription factor and geminin, negatively regulates expression of target genes in nonneuronal cells. Proc Natl Acad Sci USA. 2006;103:13074–13079. doi: 10.1073/pnas.0601915103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Kim D, Won J, Shin DW, Kang J, Kim YJ, Choi SY, Hwang M-K, Jeong B-W, Kim GS, Joe CO, Chung S-H, Song W-J. Regulation of Dyrk1A kinase activity by 14-3-3. Biochem Biophys Res Commun. 2004;323:499–504. doi: 10.1016/j.bbrc.2004.08.102. [DOI] [PubMed] [Google Scholar]
  • 37.Reeves RH, Irving NG, Moran TH, Wohn A, Kitt C, Sisodia SS, Schmidt C, Bronson RT, Davisson MT. A mouse model for Down syndrome exhibits learning and behaviour deficits. Nat Genet. 1995;11:177–184. doi: 10.1038/ng1095-177. [DOI] [PubMed] [Google Scholar]
  • 38.Sago H, Carlson EJ, Smith DJ, Kilbridge J, Rubin EM, Mobley WC, Epstein CJ, Huang TT. Ts1Cje, a partial trisomy 16 mouse model for Down syndrome, exhibits learning and behavioral abnormalities. Proc Natl Acad Sci USA. 1998;95:6256–6261. doi: 10.1073/pnas.95.11.6256. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Siarey RJ, Stoll J, Rapoport SI, Galdzicki Z. Altered long-term potentiation in the young and old Ts65Dn mouse, a model for Down syndrome. Neuropharmacology. 1997;36:1549–1554. doi: 10.1016/S0028-3908(97)00157-3. [DOI] [PubMed] [Google Scholar]
  • 40.Siarey RJ, Villar AJ, Epstein CJ, Galdzicki Z. Abnormal synaptic plasticity in the Ts1Cje segmental trisomy 16 mouse model of Down syndrome. Neuropharmacology. 2005;49:122–128. doi: 10.1016/j.neuropharm.2005.02.012. [DOI] [PubMed] [Google Scholar]
  • 41.Belichenko PV, Masliah E, Kleschevnikov AM, Villar AJ, Epstein CJ, Salehi A, Mobley WC. Synaptic structural abnormalities in the Ts65Dn mouse model of Down syndrome. J Comp Neurol. 2004;480:281–298. doi: 10.1002/cne.20337. [DOI] [PubMed] [Google Scholar]
  • 42.Belichenko PV, Kleschevnikov AM, Salehi A, Epstein CJ, Mobley WC. Synaptic and cognitive abnormalities in mouse models of Down syndrome: exploring genotype-phenotype relationships. J Comp Neurol. 2007;504:329–345. doi: 10.1002/cne.21433. [DOI] [PubMed] [Google Scholar]
  • 43.Belichenko NP, Belichenko PV, Kleschevnikov AM, Salehi A, Reeves RH, Mobley WC. The “Down syndrome critical region” is sufficient in the mouse model to confer behavioral, neurophysiological, and synaptic phenotypes characteristic of Down syndrome. J Neurosci. 2009;29:5938–5948. doi: 10.1523/JNEUROSCI.1547-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Olson LE, Roper RJ, Sengstaken CL, Peterson EA, Aquino V, Galdzicki Z, Siarey R, Pletnikov M, Moran TH, Reeves RH. Trisomy for the Down syndrome ‘critical region’ is necessary but not sufficient for brain phenotypes of trisomic mice. Hum Mol Genet. 2007;16:774–782. doi: 10.1093/hmg/ddm022. [DOI] [PubMed] [Google Scholar]
  • 45.Benavides-Piccione R, Dierssen M, Ballesteros-Yanez I, Martinez de Lagran M, Arbones ML, Fotaki V, DeFelipe J, Elston GN. Alterations in the phenotype of neocortical pyramidal cells in the Dyrk1A+/– mouse. Neurobiol Dis. 2005;20:115–122. doi: 10.1016/j.nbd.2005.02.004. [DOI] [PubMed] [Google Scholar]
  • 46.Fotaki V, Dierssen M, Alcantara S, Martinez S, Marti E, Casas C, Visa J, Soriano E, Estivill X, Arbones ML. Dyrk1A haploinsufficiency affects viability and causes developmental delay and abnormal brain morphology in mice. Mol Cell Biol. 2002;22:6636–6647. doi: 10.1128/MCB.22.18.6636-6647.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Moller RS, Kubart S, Hoeltzenbein M, Heye B, Vogel I, Hansen CP, Menzel C, Ullmann R, Tommerup N, Ropers HH, Tumer Z, Kalscheuer VM. Truncation of the Down syndrome candidate gene DYRK1A in two unrelated patients with microcephaly. Am J Hum Genet. 2008;82:1165–1170. doi: 10.1016/j.ajhg.2008.03.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Hanger DP, Brion JP, Gallo JM, Cairns NJ, Luthert PJ, Anderton BH. Tau in Alzheimer’s disease and Down’s syndrome is insoluble and abnormally phosphorylated. Biochem J. 1991;275:99–104. doi: 10.1042/bj2750099. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Masters CL, Simms G, Weinman NA, Multhaup G, McDonald BL, Beyreuther K. Amyloid plaque core protein in Alzheimer disease and Down syndrome. Proc Natl Acad Sci USA. 1985;82:4245–4249. doi: 10.1073/pnas.82.12.4245. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Park J, Yang EJ, Yoon JH, Chung KC. Dyrk1A overexpression in immortalized hippocampal cells produces the neuropathological features of Down syndrome. Mol Cell Neurosci. 2007;36:270–279. doi: 10.1016/j.mcn.2007.07.007. [DOI] [PubMed] [Google Scholar]
  • 51.Ryoo S-R, Cho H-J, Lee H-W, Jeong HK, Radnaabazar C, Kim Y-S, Kim M-J, Son M-Y, Seo H, Chung S-H, Song W-J. 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. 2008;104:1333–1344. doi: 10.1111/j.1471-4159.2007.05075.x. [DOI] [PubMed] [Google Scholar]
  • 52.Ferrer I, Barrachina M, Puig B, Martinez de Lagran M, Marti E, Avila J, Dierssen M. Constitutive Dyrk1A is abnormally expressed in Alzheimer disease, Down syndrome, Pick disease, and related transgenic models. Neurobiol Dis. 2005;20:392–400. doi: 10.1016/j.nbd.2005.03.020. [DOI] [PubMed] [Google Scholar]
  • 53.Ryoo S-R, Jeong HK, Radnaabazar C, Yoo J-J, Cho H-J, Lee H-W, Kim I-S, Cheon Y-H, Ahn YS, Chung S-H, Song W-J. DYRK1A-mediated hyperphosphorylation of Tau. A functional link between Down syndrome and Alzheimer disease. J Biol Chem. 2007;282:34850–34857. doi: 10.1074/jbc.M707358200. [DOI] [PubMed] [Google Scholar]
  • 54.Wegiel J, Dowjat K, Kaczmarski W, Kuchna I, Nowicki K, Frackowiak J, Mazur Kolecka B, Wegiel J, Silverman WP, Reisberg B, Deleon M, Wisniewski T, Gong CX, Liu F, Adayev T, Chen-Hwang MC, Hwang YW. The role of overexpressed DYRK1A protein in the early onset of neurofibrillary degeneration in Down syndrome. Acta Neuropathol. 2008;116:391–407. doi: 10.1007/s00401-008-0419-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Patterson D. The causes of Down syndrome. Sci Am. 1987;257:52–57. doi: 10.1038/scientificamerican0887-52. [DOI] [PubMed] [Google Scholar]
  • 56.Pulsifer MB. The neuropsychology of mental retardation. J Int Neuropsychol Soc. 1996;2:159–176. doi: 10.1017/S1355617700001016. [DOI] [PubMed] [Google Scholar]
  • 57.Mann DM, Esiri MM. The pattern of acquisition of plaques and tangles in the brains of patients under 50 years of age with Down’s syndrome. J Neurol Sci. 1989;89:169–179. doi: 10.1016/0022-510X(89)90019-1. [DOI] [PubMed] [Google Scholar]
  • 58.Bianchetti A, Trabucchi M, Cipriani G. Aggressive behaviour associated with donepezil treatment: a case report. Int J Geriatr Psychiatry. 2003;18:657–658. doi: 10.1002/gps.824. [DOI] [PubMed] [Google Scholar]
  • 59.Lott IT, Osann K, Doran E, Nelson L. Down syndrome and Alzheimer disease: response to donepezil. Arch Neurol. 2002;59:1133–1136. doi: 10.1001/archneur.59.7.1133. [DOI] [PubMed] [Google Scholar]
  • 60.Lobaugh NJ, Karaskov V, Rombough V, Rovet J, Bryson S, Greenbaum R, Haslam RH, Koren G. Piracetam therapy does not enhance cognitive functioning in children with down syndrome. Arch Pediatr Adolesc Med. 2001;155:442–448. doi: 10.1001/archpedi.155.4.442. [DOI] [PubMed] [Google Scholar]
  • 61.Fernandez F, Morishita W, Zuniga E, Nguyen J, Blank M, Malenka RC, Garner CC. Pharmacotherapy for cognitive impairment in a mouse model of Down syndrome. Nat Neurosci. 2007;10:411–413. doi: 10.1038/nn1860. [DOI] [PubMed] [Google Scholar]
  • 62.Costa AC, Scott-McKean JJ, Stasko MR. 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. 2008;33:1624–1632. doi: 10.1038/sj.npp.1301535. [DOI] [PubMed] [Google Scholar]
  • 63.Guedj F, Sebrie C, Rivals I, Ledru A, Paly E, Bizot JC, Smith D, Rubin E, Gillet B, Arbones M, Delabar JM. Green tea polyphenols rescue of brain defects induced by overexpression of DYRK1A. PLoS ONE. 2009;4:e4606. doi: 10.1371/journal.pone.0004606. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Bain J, Plater L, Elliott M, Shpiro N, Hastie CJ, McLauchlan H, Klevernic I, Arthur JS, Alessi DR, Cohen P. The selectivity of protein kinase inhibitors: a further update. Biochem J. 2007;408:297–315. doi: 10.1042/BJ20070797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Ortiz-Abalia J, Sahun I, Altafaj X, Andreu N, Estivill X, Dierssen M, Fillat C. Targeting Dyrk1A with AAVshRNA attenuates motor alterations in TgDyrk1A, a mouse model of Down syndrome. Am J Hum Genet. 2008;83:479–488. doi: 10.1016/j.ajhg.2008.09.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Kim ND, Yoon J, Kim JH, Lee JT, Chon YS, Hwang M-K, Ha I, Song W-J. Putative therapeutic agents for the learning and memory deficits of people with Down syndrome. Bioorg Med Chem Lett. 2006;16:3772–3776. doi: 10.1016/j.bmcl.2006.04.042. [DOI] [PubMed] [Google Scholar]
  • 67.Koo KA, Kim ND, Chon YS, Jung M-S, Lee B-J, Kim JH, Song W-J. QSAR analysis of pyrazolidine-3,5-diones derivatives as Dyrk1A inhibitors. Bioorg Med Chem Lett. 2009;19:2324–2328. doi: 10.1016/j.bmcl.2009.02.062. [DOI] [PubMed] [Google Scholar]
  • 68.Kuhn C, Frank D, Will R, Jaschinski C, Frauen R, Katus HA, Frey N. DYRK1A is a novel negative regulator of cardiomyocyte hypertrophy. J Biol Chem. 2009;284:17320–17327. doi: 10.1074/jbc.M109.006759. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Baek KH, Zaslavsky A, Lynch RC, Britt C, Okada Y, Siarey RJ, Lensch MW, Park IH, Yoon SS, Minami T, Korenberg JR, Folkman J, Daley GQ, Aird WC, Galdzicki Z, Ryeom S. Down’s syndrome suppression of tumour growth and the role of the calcineurin inhibitor DSCR1. Nature. 2009;459:1126–1130. doi: 10.1038/nature08062. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Cellular and Molecular Life Sciences: CMLS are provided here courtesy of Springer

RESOURCES