Abstract
Genomically identical cells have long been assumed to comprise the human brain, with post-genomic mechanisms giving rise to its enormous diversity, complexity, and disease susceptibility. However, the identification of neural cells containing somatically generated mosaic aneuploidy – loss and/or gain of chromosomes from a euploid complement – and other genomic variations including LINE1 retrotransposons and regional patterns of DNA content variation (DCV), demonstrate that the brain is genomically heterogeneous. The precise phenotypes and functions produced by genomic mosaicism are not well understood, although the effects of constitutive aberrations, as observed in Down syndrome, implicate roles for defined mosaic genomes relevant to cellular survival, differentiation potential, stem cell biology, and brain organization. Here we discuss genomic mosaicism as a feature of the normal brain as well as a possible factor in the weak or complex genetic linkages observed for many of the most common forms of neurological and psychiatric diseases.
Keywords: Aneuploidy, genome, mosaicism, somatic mutation, CNS, CNV
Introduction
Aneuploidy is a gain (hyperploidy) or loss (hypoploidy) of chromosomes such that the resulting chromosome number is not an exact multiple of the haploid complement. A related term, aneusomy, reflects specific chromosome gains (hypersomy) or loss (hyposomy) in a cell, although the full karyotype for that cell may be unknown relative to the germline chromosomal complement. Aneuploidies and aneusomies within an organism can be defined as either constitutive, meaning that changes begin in the germline or early embryogenesis, resulting in a conserved change in virtually all cells of an organism; or mosaic, which indicates somatic changes in individual cells that result in mixed aneuploid and euploid forms with varied prevalence throughout an organism. There are several well-known pathophysiological chromosomal disorders including Down (trisomy 21), Edwards (trisomy 18), and Patau (trisomy 13) syndromes, which are most commonly constitutive in >95% of cases [1-5], along with sex chromosome aneuploidies like Klinefelter’s (XXY) and Turner’s (monosomy of X) syndromes that also result in abnormal development and behavior [6-10]. Mosaic disorders affecting the brain have also been described, such as mosaic variegated aneuploidy (MVA) [11-15].
While such chromosomal aberrations have been long associated with neurogenetic disorders, chromosomal aneuploidies or aneusomies are also known to be a normal feature of the brain, manifesting as complex mosaics [16-28]. In the central nervous system (CNS), mosaic aneuploidies were first identified in the cerebral cortex of normal developing mice [23], a result that has been extended throughout the neuraxis and to all vertebrate species thus far examined [21, 22, 25], including non-diseased humans [19, 24, 25, 27, 28]. Moreover, these changes have been a harbinger for other genomic alterations, generally referred to as DNA content variation (DCV) [29] 2010). Here we discuss genomic mosaicism in the non-diseased brain, and how it may contribute to human brain diseases.
1. Genomic diversity in cells of the normal brain: Mosaic aneuploidy and DNA content variation (DCV)
1.1 Detection techniques
As early as 1902, Theodor Boveri identified chromosome aberrations in cancerous tumors, demonstrating the existence of living, aneuploid cells [30]. The simplest evaluations of chromosome numbers merely count chromosomes in metaphase spreads, when the condensed state of the chromatids allows for visualization, as well as identification of balanced and unbalanced translocations by Giemsa staining [31]. Despite the simplicity of this assay, it is notable that the correct human complement of chromosomes was not established until 1956 [32], some three years after report of the double helix [33], underscoring ambiguities that are associated with chromosome counts. A definitive modern technique called spectral karyotyping, or SKY, relies on the hybridization of genomic fragments labeled with distinct fluorochromes to the metaphase spreads of single cells and the subsequent identification of each chromosome pair or sex chromosomes [34] (Figure 1A). These strategies require condensed chromosomes, and as such cannot be definitively used on interphase or non-mitotic cells. Fluorescent in situ hybridization (FISH) also employs hybridization of a probe against a defined but limited chromosomal region (“point probes”), which can be used to assess aneusomies in single interphase cells using a fluorescent or enzymatic readout (Figure 1B). Multicolor FISH allows for simultaneous evaluation of several chromosomes or different regions along a single chromosome, including quantification of FISH signal intensity [35]. However, there are technical limitations that can lead to false-positive and false-negative probe hybridization, which require careful controls to identify true aneuploidy versus artifactual hybridization, such as pairing of chromosome homologs that may lead to the incorrect interpretation of a “pseudo monosomy” [27]. A modification of point probe FISH is interphase chromosome-specific multicolor banding (ICS-MCB) wherein a set of specific paints derived from microdissected chromosomes labels the target chromosome with a distinct spectral pattern for the simultaneous visualization of several regions of the chromosome [36, 37]. This technique has not been widely used and may depend on the cell type and/or age of the interrogated chromatin. An independent technique for chromosomal copy number analysis is comparative genomic hybridization (CGH) and array CGH [38, 39]. CGH requires the hybridization of test genomic samples to a representation of a standardized genome, and allows for copy number analyses from tissue samples or prenatal cytogenetic samples. Previously, the requirement of a relatively large, genomically homogenous set of cells limited the use of CGH in identifying mosaic aneuploidy. While Ballif and colleagues reported the detection of mosaicism even at levels of 10-20% [40], its effectiveness in CNS samples remains to be determined.
Figure 1.
Schematic of genomic mosaicism analysis techniques. A, Cells from cycling populations may be arrested in metaphase for either chromosome spread enumeration by counting DAPI-stained chromosomes (bottom left), or full karyotype analysis by SKY (bottom right). B, Non-cycling or interphase cells are hybridized with chromosome-specific FISH probes (e.g., the chromosome 8 and 16 point probes shown here in red and green, respectively). Euploid cells, disomic for both chromosomes, would display 2 dots of each; here, both nuclei are disomic for chromosome 8, while the nucleus on the left is monosomic for chromosome 16 and the nucleus on the right is trisomic for chromosome 16. C, Isolated cells and nuclei are stained to saturation with dyes like the DNA-intercalating dye Propidium Iodide or DRAQ5 for flow cytometric analysis. The prominent peak of the resulting DNA content histogram contains cells in the G0/G1 phase of the cell cycle (2N DNA content); S phase (2<N<4) and G2/M (4N) phase are distinguishable on the linear scale of the x-axis. D, Hetergeneous DNA content histogram from human frontal cortical nuclei (green, red and blue are separate individuals) stained with propidium iodide, showing broad bases and right-hand shoulders. Chicken erythrocyte nuclei (CEN) were included as an internal reference standard and control. E, Overlay of representative lymphocyte (green), cerebellar (red) and cortical (blue) histograms displaying an area of increased DCV unique to the cortical sample. (Adapted with permission from Peterson et al., 2012; Westra et al., 2010.)
Single cell approaches that are currently in development will help to lower the detection threshold. The genome from single cells isolated by laser microdissection, flow cytometry, or other techniques could be amplified in a uniform and unbiased manner (e.g., using multiple displacement amplification (MDA) [41]) for analysis by single-cell CGH or quantitative PCR for target genomic regions. Even more definitively, the resulting amplicons from single-cell MDA could serve as a template for genomic sequencing, an approach being pursued for cancer cells [42, 43], as well as partial sequencing from neurons [44]. The promise of these techniques is currently tempered by a range of factors including use of adequate control genomes, the current low throughput of the technique that is critical in view of the one trillion cells that make up the human brain, and sufficient information storage limitations for the terabytes of data produced by whole-genome sequencing.
A distinct approach to assessing genomic uniformity is DNA flow cytometry that has a long history of identifying cells with varying DNA content associated with phases of the cell cycle [45, 46]. The highly integrated and physically connected nature of the brain (e.g., its synaptic neuropil) makes analyses of single cells difficult and incomplete, thus limiting prior flow cytometry efforts for studying the brain. Modifications of this approach to interrogate isolated nuclei rather than intact cells from the brain for DNA content (Figure 1C) has identified brain cell populations with a surprising range of DNA content (Figure 1D). This was manifested as an overall increase in DNA content within cerebral cortical neurons compared to cerebellar neurons from the same individual, demonstrating the pervasive existence of normal human brain cells having DNA content variation (DCV) (Figure 1E) [29]. DCV in the frontal cortex averages a gain of 250Mb, with NeuN-positive neurons showing significant increases compared to non-neuronal nuclei. Importantly, DCV appears to encompass myriad forms of mosaic aneuploidy that exist in both the cerebral cortex and cerebellum [24, 25, 27, 28]. By contrast, DCV also appears to be distinct from aneuploidy because of the expanded DNA content histograms in the cerebral cortex that are less prominent in the cerebellum, suggesting an independent mechanism for increased DNA content.
These technical approaches, along with others in development, have allowed assessments of single brain cells, demonstrating genomic mosaicism amongst cells of the brain – and likely other tissues and cells, including stem cell lines [47, 48] – thus redefining the genomic organization of the brain from homogenously uniform to a complex genomic mosaic. These data underscore a need to consider individual genomes in cellular function in the normal and diseased brain, as well as the effects of identified genes operating in varied genomic surroundings.
1.2 Mosaic aneuploidy in the non-diseased brain
The first report of widespread genomic mosaicism came from studies of aneuploidy in mice, which revealed that approximately 33% of proliferating cerebral cortical neural progenitor cells (NPCs), isolated from the ventricular zone of the embryonic brain [23], were aneuploid. A range of other neurogenic regions generate aneuploid cells, including cerebellar NPCs that represent ~15% of mitotic cells at postnatal day (P) P0 and ~21% at P7 [17, 25]. This somatically derived form of genomic variation is characterized by the apparently stochastic loss or gain of all chromosomes, creating a genomic mosaic that displays a predominance of hypoploidy over hyperploidy [23]. During periods of cell division, mosaic aneuploidy in NPCs results from chromosomal segregation defects (lagging chromosomes, non-disjunction and supernumerary centrosomes) during mitosis [26, 49]. Some mosaically aneuploid cells remain capable of differentiating into neuronal and glial lineages [17], and can survive into the adult brain [23], where they can be integrated into active neural circuitry [18]. In the mosaic landscape of the CNS, the genomic diversity of aneuploid cells, and the subsequent differences in gene expression profiles [17, 50-52] suggest that a great deal of cellular variability and diversity exists without negatively impacting the high functionality of the system.
In the human CNS, the same trends have been reported: the genomic variation caused by aneuploidy in the developing brain reaches 30-35%, while most other tissues display low, albeit detectable, levels of aneuploidy [20, 28]. While the total amount of aneuploidy in the mature human brain remains unknown, reflecting both the size of the brain and the limitations of current evaluation techniques, several studies provide evidence that a significant population of aneuploid cells is also present in the adult human brain. Rehen and colleagues tracked chromosome 21 in neurons and non-neuronal cells from the frontal cortex and hippocampus of non-diseased human brains (aged 2 through 86), using dual-locus hybridization that combined a chromosome paint with a FISH point probe for increased specificity, reporting a rate of ~4% aneusomy for chr 21, with monosomy more frequent than trisomy [24]. Several subsequent studies have attempted to give a more complete analysis of the aneuploidy rate in non-diseased adult human brains by analyzing several chromosomes, including 21, using mixes of multiple enumeration probes or ICS-MCB. These studies reported lower frequencies of aneuploidy (0.1-0.8% on average) for an estimated total aneuploidy level of 10% [16, 53, 54]. It is critical to note that all of these studies suffer from: 1) an inability to objectively identify trisomy – subjectivity is inherent in scoring ambiguous hybridization patterns; 2) an inability to precisely compare the same cell types and brain regions between different individuals; and 3) severe limitations of sample size, as interrogation of even 10,000 cells represents less than 0.000001% of the more than 1 trillion cells in the human brain. The differences in total observed aneuploidy levels may be complicated by the precision of the techniques used to evaluate aneuploidy, discussed above. What can be concluded is that developmental aneuploidy amongst NPCs is robust, while aneusomies in adult brains unambiguously exist but at levels that require further clarification to determine aneusomy rates for all chromosomes simultaneously.
1.3 Functions of neural mosaic aneuploidy
The functional significance of neural mosaic aneuploidy is beginning to emerge. Aneuploidy has clear cellular and organismal consequences, as seen in analyses of genomically unstable cancers and constitutively aneuploid diseases like Down syndrome [55-57]. As noted above, aneuploidy in cells is known to affect gene expression compared to euploid counterparts in a range of organisms, from yeast to mammals [17, 50-52, 58-60]. Aneuploidy can affect a range of cellular processes including survival, proliferation potential, and protein imbalances [58, 61-64]. The integration of adult aneuploid neurons into the circuitry of the normal brain [18] therefore suggests the potential of these neurons to influence normal brain functions. However, proven consequences of mosaic aneuploidy in the CNS have been difficult to establish because of the difficulty in identifying the mechanistic link between a specific karyotype and an identified function in a living cell. Not only have aneuploidies been overwhelmingly studied on fixed, non-living cells therefore precluding functional studies, but the loss or gain of a chromosome affects the expression of all genes and regulatory regions therein, creating an intricate web of interconnected consequences. Use of GFP reporters integrated into a defined chromosome have enabled gene expression analyses on cells with defined aneusomy (with the associated loss of GFP) vs. normal cells, and this approach indicates that aneuploidies can alter gene expression profiles within seemingly homogenous populations of neural cells [17]. In light of these observations, it would be surprising if mosaic aneuploid cells within the brain did not have functional consequences. Moreover, it is notable that aneuploidies are species-specific by virtue of unique chromosome organization and number that in part define a species. Thus, gains and/or losses of identified chromosomes within one species would be expected to have non-identical, albeit possibly overlapping effects on the most closely related chromosome from another species (i.e., based on the degree of synteny between chromosomes of the compared species).
Recent studies on the functional consequences of neural aneuploidy identified effects of mosaic aneuploidy on developing NPCs [65], wherein identified forms of aneuploid cells are differentially eliminated by caspase-mediated programmed cell death (PCD). These data support a mechanism of cell selection based on somatically generated genomic variation produced by aneuploidy [65, 66]. Within the developing murine brain, PCD eliminates cells to control cerebral cortical shape and size [67-71]. Blocking PCD, mediated by either the effector caspase-3 or the initiator caspase-9, produces severe exencephaly, an expanded ventricular zone, NPC hyperplasia, and death [72-78]. Remarkably, suppression of PCD by genetic deletion of either caspase-3 or caspase-9, or through pharmacological inhibition of caspases, results in a concomitant increase in total aneuploidy in mitotic and post-mitotic cerebral cortical cells, as well as in “extreme” forms of aneuploidy (operationally defined as cells with a gain or loss of greater than 5 chromosomes; nullisomies, where both copies of the same chromosome are lost; and the coincident losses and gains of chromosomes) [65]. Interestingly, comparatively mild forms of aneuploidy (cells that gained or lost fewer than 5 chromosomes) remained relatively unchanged, indicating preferential effects of PCD in removing extreme forms of neural cell aneuploidy while maintaining mildly aneuploid and euploid populations. Changes in mosaic aneuploidy levels and forms produced by reducing PCD likely contribute to the phenotype, including early postnatal lethality, of caspase-null mutants [72, 73]. These data provide the first evidence of functional consequences for distinct forms of mosaic aneuploidy during CNS development, as extreme forms of aneuploidy – hypoploidy and hyperploidy – are eliminated by PCD, contrasting with survival of mildly aneuploid and euploid cells. Mosaic aneuploidies create – by definition – genomic diversity amongst populations of brain cells, which would be expected to have functional consequences based on changes in gene expression produced by aneuploidy [17, 58, 60-63]. Maintenance of seemingly neutral or beneficial aneuploidies [79] and euploid cells may therefore be the end-result of selective pressures that are a consequence of mosaic, somatic genomic alterations during CNS development [65]. The concept of beneficial aneuploidies is particularly intriguing, in that the loss or gain of the chromosome may provide some ability to the cell that makes it more “fit” than other cells from the same organism (for example, increased stress resistance or an enhanced functional capacity within a neural network). The difficulty of linking an aneuploid event with a particular function, as mentioned above, has thus far limited the identification of a specifically beneficial set of aneuploidies, but it is important to note that there is a fine line between fitness and detriment – a cell with decreased response to stress signals or cell death triggers, may share these traits with a cancer cell. Studies to identify functional consequences of particular aneuploidies will be essential in understanding roles for aneuploid cells in the CNS, and the selective pressures that may allow maintenance of mildly aneuploid states.
The existence and prevalence of mosaic aneuploidies is proof of a normally mutable genome, and it would therefore be anticipated that other forms of genomic change could exist in cells of the brain. Following the demonstration of mosaic aneuploidies, the identification of possibly amplified repeat elements like LINE1 retrotransposable elements was reported [80, 81] that have been proposed to “jump” amongst neurons. However, this mechanism may not occur sufficiently to account for increased diversity, at least within the human cerebral cortex [29, 44]. More broadly, the identification of DCV that is best manifested as DNA gains within the human cerebral cortex, encompasses aneuploidies, possible LINE retrotransposons, as well as other genomic changes. The actual origins of DCV are not known but could involve a range of reported structural variants that include not only aneuploidy, but also somatic versions of copy number variation (CNV) [82] and other alterations to the genome, in view of the evidence for genomic, rather than extragenomic, origins of the increases in DNA content [29].
1.4 Mosaic aneuploidy in pluripotent stem cell lines
Mosaic aneuploidy has been clearly demonstrated in NPCs in the CNS, but the effective study of its functional consequences is limited, as noted above, by available experimental paradigms. Human embryonic stem cells (hESCs) and induced pluripotent stem cells (iPSCs) offer an attractive in vitro system to examine cellular processes that could be affected by mosaic aneuploidy, including differentiation, development, and neurological models. This approach has been supported by the observation that stem cells also show genomic heterogeneity produced by aneuploidy or other genomic alterations like CNVs [47, 48, 83-87]. Culture-induced aneuploidies have been observed in hESCs [88, 89] – in particular, gains of chromosomes 12, 17, 1, or X have been reported, which may arise by imparting a selective growth or survival advantage to cells with these karyotypes; cells with these recurrent gains can overwhelm the culture, leading to a clonal constitutively aneuploid cell population (i.e., all cells of the culture exhibit the same aberrant karyotype) [79]. In contrast, consistent with observations on mouse ES cells [87], a recent study found that ~18-35% of cells within a given hESC line show mosaic aneuploidy, suggesting that the stochastic loss and/or gain of chromosomes is an inherent characteristic of stem cell biology [48], and this is also consistent with results from analyzing NPCs. In this study, six commercially-available hESC lines and an iPSC line derived from fibroblasts showed significant levels of mosaic aneuploidy, independent of passage number and cell culture conditions (varied media, supplements and substrates, and investigators) [48]. It is important to emphasize the difference between the stochastic generation of mosaic aneuploidy versus clonal constitutively aneuploid karyotypes in long-term culture, particularly with respect to stem cell usage as a therapeutic. Mosaic aneuploidy mimics the genomic variability observed in vivo and likely contributes to the normal phenotypic heterogeneity of gene expression patterns [90, 91]. Devalle and colleagues suggest that mosaicism in stem cell culture may be well tolerated as cells can respond divergently to a range of stimuli, but the low levels and the random nature of this aneuploidy does not impart a selective clonal advantage to cells. Conversely, inundation of a culture by cells with clonal constitutive changes, arising from adaptations to stressful or unhealthful environments [79], represents a challenge for stem cell safety and usage in vivo. Such changes have been correlated with cancers, such as the loss of chromosome 10, gain of chromosome 7 or chromosome 1p and 19q deletions in human gliomas [55, 57, 92-95], as well as the identification of cancer related genes on the most common aneuploid chromosomes in hESC culture [88, 96, 97]. The carcinogenic risks that mosaically aneuploid hESCs pose to transplantation therapies – as well as within mosaic populations in the developing and adult brain – remain to be determined, but raise the formal possibility of aneuploid progenitor cell populations as a source of cancer stem cells in the brain, as well as other tissues. These issues deserve further analyses, especially where transplantation is designed to incorporate cells for the lifetime of an individual, as would be desired for neurons. The effects of transplanting stem cells or their derivatives having identifiable genomic mosaicism can be gleaned from studies linking neuropathological effects to aneuploidy, as discussed next.
2. Linking genomic mosaicism and brain diseases
2.1 Down syndrome
Three human constitutive autosomal trisomies are compatible with live birth – trisomy 13 (Patau’s syndrome) [98], 18 (Edward’s syndrome) [99], and trisomy 21 (Down syndrome (DS)) [100-102]. The severity of these chromosomal disorders suggests an inverse correlation between the amount of genetic material gained vs. the severity of the resultant phenotype and organismal fitness [52]. The most common chromosomal disorder is Down syndrome, in which the smallest autosome (21) is duplicated to produce a trisomy. Trisomy of chromosome 21 results in abnormal brain development (along with non-CNS malformations) including reduced brain size with fewer neurons, leading to moderate to severe mental retardation, and early-onset Alzheimer’s disease (AD)-like clinical signs and neuropathology [103, 104]. These traits can at least in part be attributed to the triplication of 33 genes in the DS critical region (DSCR) of chromosome 21, leading to increased gene dosage and imbalances in affected genetic regulatory circuits [105-107]. The DSCR triplication as typically identified from peripheral blood or other non-CNS tissues may occur in several ways: complete trisomy, which 95% of DS patients have [101, 102, 108]; mosaic aneuploidy of intermixed cells of disomic and trisomic chromosome 21, accounting for ~2% of DS patients [109]; rarer partial trisomy of regions greater than 5 Mb; and microtrisomies of 3-5 Mb regions generated by unequal meiotic crossovers, the prevalence of which is not known because of detection difficulty by standard cytogenetics [56].
Within the DSCR, studies have identified several candidate genes that may impact the neurological and cognitive dysfunctions of DS (Table 1), from developmental brain defects to the early onset of AD-like neurodegeneration: Amyloid precursor protein (APP), Superoxide dismutase 1 (SOD1) [106], Regulator of calcineurin (RCAN1 or DSCR1) [107, 110, 111], and Dual-specificity tyrosine-(Y)-phosphorylation regulated kinase 1A (DYRK1A) [107, 112-114]. Of these, APP is best-known as a key component of senile plaque formation in AD (discussed below). SOD1 protects against the free radical superoxide anion O2− by reducing it to hydrogen peroxide (H2O2) and O2 in non-diseased individuals. The over-expression of SOD1 in DS is thought to create an imbalance in the free radical detoxication system leading to oxidative stress (OS) [115-117], which in turn promotes DNA damage and chromosomal nondisjunction that may lead to aneuploidy, and increased cell death contributing to abnormal brain development and mental retardation [114, 118, 119].
Table 1. Genes and chromosomal regions associated with neuropsychiatry disorders disorders that: may be altered by genomic mosaicism.
Disease | Locus | Gene | Protein | Role of region |
---|---|---|---|---|
Down
Syndrome |
21q21.3 | APP | Amyloid precursor protein | Senile plaque formation |
21q22.11 | SOD1 | Superoxide dismutase 1 | Oxidative stress response | |
21q22.12 | RCAN1/ DSCR1 |
Regulator of calcineurin 1 | Oxidative stress response | |
21q22.13 | DYRK1A | Dual-specificity tyrosine-(Y)- phosphorylation regulated kinase 1A |
Protein phosphorylation, including Tau | |
Alzheimer’s
Disease |
21q21.3 | APP | Amyloid precursor protein | Senile plaque formation |
14q24.3 | PSEN1 | Presenilin 1 | Regulator of gamma-secretase activity | |
1q31- q42 |
PSEN2 | Presenilin 2 | Regulator of gamma-secretase activity | |
17q21.1 | TAU | Microtubule-associated protein tau |
Neurofibrillary tangle formation | |
Schizophrenia | 15q11.2 | UBE3A; NIPA1 |
Ubiquitin protein ligase 3A; Non imprinted in Prader-Willi syndrome 1 |
Region implicated in Angelman syndrome or Prader-Willi syndrome |
1q21.1 | Deletion or duplication | |||
15q13.3 | Microdeletion | |||
16p11.2 | Duplication | |||
16p13.1 | Duplication | |||
22q11.2 | Microdeletion resulting in DiGeorge syndrome |
|||
Autism | 15q11- 13 |
GABRB3; GABRA5 |
ɣ -amino butyric acid (GABA) receptor-type A subunits 3 and 5 |
GABAergic neuron development |
Xp22.11- p21.2 |
IL1RAPL1 | Interleukin 1 receptor accessory protein-like 1 |
Hippocampal memory system | |
Xq13.1 | NLGN3 | Neuroligin 3 | Synapse formation and remodeling | |
Xp22.33 | NLGN4X | Neuroligin 4, X-linked | Synapse formation and remodeling | |
MVA | 15q15 | BUB1B | Budding uninhibited by benzimidazoles 1 homolog beta |
Spindle checkpoint regulation |
RCAN1, or DSCR1, is thought to play a role in brain development, long-term potentiation, and memory, and is significantly elevated in both DS and AD [120-122]. Like SOD1, RCAN1 is thought to play a role in the OS response: the short-term induction of one highly-expressed neuronal isoform (RCAN1.1L) protects the cell from OS-induced apoptosis, while its long-term expression or over-expression, as in DS, promotes neuronal susceptibility to OS-induced apoptosis [111, 123-125]. In addition to the OS response and neuronal death, RCAN1 has other functions that may promote the AD neuropathology seen in DS adults. Brain-expressed RCAN1 isoforms can both inhibit calcineurin [110, 122], a serine/threonine phosphatase, and induce glycogen synthase kinase (GSK)-3β activity [126]. This is relevant as these enzymes regulate the phosphorylation of microtubule-associated protein Tau, suggesting that RCAN1 contributes to Tau hyperphosphorylation and the formation of AD neurofibrillary tangles [126, 127]. Intriguingly, the over-expression of RCAN1 can also be induced by amyloid-β1-42, the cleaved peptide product that comprises AD plaques [120], providing another link between this DS gene and AD.
The serine/threonine kinase DYRK1A is also overexpressed in the DS brain, and has numerous links to DS phenotypes. DYRK1A protein modulates several transcription factors involved in an array of functions during development and adulthood: postembryonic neurogenesis, dendritic development, synaptic plasticity and memory [112, 128], while over-expression leads to significant impairment of spatial learning and memory [107, 114, 129]. The combined overexpression of DYRK1A and RCAN1, mentioned above, has a cooperative effect on the dysregulation of developmental transcription factor NFAT, contributing to DS neural deficits [130]. Increased DYRK1A kinase activity also has direct, gene dosage-dependent effects on neurofibrillary tangle formation in neurodegeneration, as DYRK1A directly phosphorylates Tau at 11 sites and primes Tau for further phosphorylation by GSK-3β, enhancing Tau self-aggregation [112, 113, 131, 132]. Compounding the effects of DYRK1A neurofibrillary degeneration in DS, as several recent studies have reported, DYRK1A-mediates disruption of Tau RNA alternative splicing. Inclusion or exclusion of Tau exon 10 gives rise to two isoforms, the imbalance of which contributes to neurofibrillary degeneration; normal DYRK1A phosphorylation promotes exon 10 splicing [133], but upregulated DYRK1A activity suppresses exon 10 inclusion and disrupts the isoform ratio [134].
The ties between these genes, amongst others on chromosome 21, and the phenotypic traits of Down syndrome illustrate the gene expression and phenotypic consequences of genomic dosage imbalances caused by aneuploidy, and potentially altered by other forms of somatic genomic mosaicism. While functional analyses of chromosome 21 genes in DS continue, the effects of trisomy 21 also have intriguing implications for understanding AD.
2.2 Alzheimer’s disease (AD)
AD is the most common dementia, impacting an estimated 5.4 million people in the United States alone (1 in 8 people over the age of 65) [135]. AD neuropathology includes the accumulation of plaques composed of amyloid β, neurofibrillary tangles containing Tau, synaptic loss and neuronal death in several brain regions, including the hippocampus, and frontal and entorhinal cortices, leading to progressive cognitive decline. Familial AD makes up ~5% of all cases, with early onset (<60 years) caused by inherited autosomal dominant mutations in predominantly 3 genes (Table 1): Amyloid precursor protein (APP) on chromosome 21, which is cleaved to form the Amyloid (A) β peptide found in amyloid plaques; presenilin 1 (PSEN1) on chromosome 14, and presenilin 2 (PSEN2) on chromosome 1, which contribute to the catalytic activity of the Aβ-cleaving enzyme ɣ -secretase. Late-onset (>60 years) sporadic AD arises from a less understood set of genetic, epigenetic, and environmental risk factors [19, 136, 137], but shares the same neuropathology with familial cases.
The classic theory of AD neuropathology is the amyloid cascade hypothesis: aggregation of a 42 amino acid peptide, Aβ1-42, which results from targeted APP cleavage. This aggregation leads to plaque formation, triggers a microglial and astrocyte-mediated inflammatory response, promotes Tau neurofibrillary tangle formation and eventually leads to neuronal death and patient dementia [138, 139]. Familial mutations in APP or the PSEN genes drive this aggregation in early onset disease, while in late-onset sporadic disease, age-dependent Aβ accumulation is proposed to cross a threshold leading to disease onset. Down syndrome brains were vital in formulating this hypothesis, as Aβ was first identified in DS plaques and later found to be the same peptide isolated from AD brains [140, 141]. The Aβ precursor, APP, was mapped to chromosome 21 [142-144], creating a stronger link between the increased gene dosage in DS produced by trisomy 21 and the earlier AD onset observed in DS patients [145].
Analyses of the chromosomal complement in AD patients, searching for a link between chromosome 21 gene dosage and the disease, have yielded mixed and conflicting results. Examinations of sporadic and familial AD samples utilized restriction fragment length polymorphisms (RFLPs) to interrogate duplications and failed to observe dosage differences [146, 147]. A more recent study has shown the duplication of the APP locus in familial early-onset AD using the much more sensitive quantitative PCR as well as FISH [148]. Several studies in peripheral lymphocytes or cultured fibroblasts from both sporadic and familial AD revealed low-level trisomy 21 in AD compared to controls [149-151], as well as an increased frequency of micronucleus formation in lymphocytes, resulting from chromosome missegregation [152]. Within the brain, Iourov and colleagues report a significant increase in aneuploidy of chromosome 21 in AD brains (from 1.7% in controls to 10.7% in disease) using ICS-MCB, but puzzlingly not with FISH. The group also points out that monosomy is as prevalent as trisomy [16], consistent with reports of both chromosomal 21 monosomies and trisomies in neurons from the frontal cortex of non-diseased individuals [24], yet incongruous with evidence of APP locus gain, rather than loss, in AD [153]. Mosch and colleagues also tracked aneuploidy in the AD brain compared to the non-disease brain by chromosome 17 FISH, and report that while monosomy levels remain consistent, trisomy 17 levels are increased in the AD brain [154]. These data, the authors suggest, support a hypothesis quite distinct from the amyloid cascade – the hypothesis that aberrant cell cycle re-entry drives DNA replication and leads to AD neuropathology [155-157].
This pathogenic cell cycle hypothesis suggests that mature, fully differentiated neurons are prompted by internal and external factors to leave the senescent G0 phase and re-enter the cell cycle. This postulated reactivation of the DNA replication pathway, while contrary to classical belief, has received some support by the ectopic neuronal expression of cell cycle proteins including cyclins D and E, cyclin-dependent kinase (cdk) 4, Ki-67, and DNA polymerase [158-166]. The cyclin D/cdk4 complex regulates progression into G1, in response to mitogenic triggers, while cyclin E expression drives the transition from G1 into S phase, when DNA replication occurs [167]. Cell cycle re-entry in the AD brain is not thought to reach completion, but rather terminates after S phase [168], which predicts the existence of neurons with 4N DNA content. Westra and colleagues performed detailed DNA content flow cytometry combined with FISH analyses of neuronal and non-neuronal nuclei isolated from AD and control brains to assess the presence of tetraploid neurons that should be the result of full chromosomal duplication, with a 4N karyotype following cell cycle replication. By contrast, no 4N neurons were detected, while some non-neuronal cells did show evidence for tetraploidy [153], consistent with glial cell division. In addition to the lack of tetraploid neurons observed in AD brains, cell cycle protein expression shows aberrant subcellular localization whereby they are detected in the cytoplasm rather than the nucleus [159, 169], suggesting that if indeed cell cycle re-entry is occurring, it is not occurring in a classical manner.
A “two-hit” hypothesis has attempted to integrate dysregulation of mitogenic processes, which contribute to somatic genomic variation including aneuploidy, and oxidative stress that accompanies aging into a cohesive explanation of AD etiology [118, 168, 170, 171]. In this hypothesis, cell cycle protein expression is thought to precede the formation of plaques and neurofibrillary tangles [164], but several AD-linked gene products have roles in cell cycle events and the missegregation of sister chromatids. Both APP and the Aβ1-42 cleavage product have been identified as mitogenic in vitro [172], inducing the phosphorylation of proteins including Tau [120, 173] and disrupting the mitotic spindle leading to defective chromosome segregation [174-176]. Presenilins also mediate premature chromosome separation. These proteins localize to kinetechores and centrosomes [177], and when overexpressed or mutated, directly lead to chromosome missegregation and low-level aneuploidy [149, 178]. Aneuploidy, like that of chromosome 21, is a secondary result of AD according to this hypothesis, that would then work in a positive feedback loop to increase APP aggregation and neuronal cell death [168, 171, 179]. Oxidative stress represents the second hit of the hypothesis. OS increases with age [157, 168], precedes Aβ deposition in AD and DS [180-184], and is linked to neurodegeneration [118, 119]. Oxidative stress and mitogenic misregulation are postulated to act as simultaneous and complementary triggers for AD. OS responses may include cell cycle re-entry and activation of AD-related proteins like APP, Tau, and the PSENs, while resultant aneuploidy or related gene dosage effects drive amyloid deposition, Tau aggregation, decreased OS response, and neurodegeneration [119, 168, 185]. Genomic variation present in the normal brain preceding the onset of disease may affect the responses of cells to environmental factors like oxidative stress, making them more or less susceptible to the onset of disease. Thus, the level of reaction to the trigger would depend on the initial genomic state of the cell, making an understanding of genomic mosaicism essential for clarification of disease progression.
In all these hypotheses – the amyloid cascade, DNA replication, and the two hit hypothesis of AD etiology – key questions remain unanswered about what makes specific brain regions vulnerable to AD onset, particularly in the case of sporadic AD. Down syndrome provides the clearest example that disease-promoting increases in gene dosage can be produced by aneuploidy (trisomy 21), however the lack of clear evidence for this process in AD indicates the need to better understand the relationship between mosaic genomic alterations and disease.
2.3 Schizophrenia
Schizophrenia is a complex multigenic neuropsychiatric disorder with onset in the late teens or early twenties that is characterized by symptoms of psychosis that may include hallucinations, delusions, decreased emotional expression, and behavioral deficits. The strongest established risk factor is family history of the disease, with estimates of 70-90% heritability based on monozygotic twin and sibling studies [186-188], but the disease does not display Mendelian inheritance. While genome-wide association studies have identified common polymorphisms in schizophrenia that confer a small to moderate disease risk [189-191], the common disease-rare variant (CD-RV) model suggests that the heterogeneity of the disease arises from multiple rarer variants with higher risk [192, 193]. Six constitutive rare CNVs (both sub-chromosomal duplications and deletions) that strongly associate with the disease have been identified by microarray analyses in large cohorts of schizophrenic patients (Table 1): 1q21.1, 15q11.2, 15q13.3, 16p11.2, 16p13.11, and 22q11.2 [191, 194-196]. The 22q11 microdeletion (a 3Mb loss) is the best characterized, as it results in velocardiofacial syndrome/DiGeorge syndrome, causing cognitive deficits, and mental disorders like schizophrenia in approximately 30% of those who have the deletion [196-198]. Other rare CNVs have a much lower frequency in schizophrenia: deletions of 1q21.1 and 13q13.3 occur in 1 in 500 patients versus 1 in 5000 controls; duplications in 16p13.1 and 16p11.2 occur in 1 in 300 patients versus 1 in 1000 or 1 in 3500 controls (respectively). These CNVs are associated with approximately 10-fold increased risk of schizophrenia, but have low penetrance – calculated between 2 and 7.4% for most CNVs [199]. The phenotypic spectrum of the CNVs linked to schizophrenia remains to be thoroughly investigated, particularly in terms of the neurodevelopmental consequences of the deletions and duplications. Additionally, there are several CNVs that may link schizophrenia to other developmental disorders, such as autism, that warrant investigation.
In addition to these CNVs, several studies have reported increases in aneuploidy associated with schizophrenia. Sex chromosomal abnormalities have been observed in lymphocytes from small numbers of patients compared to controls, analyzed primarily by metaphase spreads. Karyotypes of 47, XXX and 47, XXY were most commonly associated with increased schizophrenia, with schizophrenic patients displaying aneuploidy at rates four to six times that of the general population [200, 201]. Occasional patients also had mosaicism within examined lymphocytes (45,X/46, XX; 45,X/46,XY; 46,XX/47,XXX; 46,XY/47,XXY) [19, 200]. It is important to note that aneuploidy does not seem to be a common characteristic of all schizophrenic patients, as less than 2% of schizophrenic patients also had chromosomal anomalies in peripheral lymphocytes [200]. Only Yurov and colleagues have sought to address aneuploidy levels in the brain, reporting both low-level sex chromosome and chromosome 1 mosaic aneuploidy in schizophrenia [202, 203]. In the first study, two individuals in an eight-sample cohort had chromosome X trisomy (cells with fewer than 2 signals were excluded, to rule out pseudo-monosomies), although very low numbers of brain cells were counted (200 nuclei) [202]. A more detailed study using both multicolor FISH probes against several chromosomes (including chromosome 1 and X) and ICS-MCB probes detected low-level mosaic chromosome 1 aneuploidy in 2 of 12 schizophrenic brain samples examined, but did not reveal any sex chromosome anomalies [203]. The small number of total schizophrenic brains analyzed that showed chromosomal aneuploidy speaks to the widespread heterogeneity of the disease, and illustrates the need for more comprehensive, detailed study on the role of aneuploidy in neuropsychiatric disorders.
2.4 Autism
Autism is another complex neurodevelopmental disorder broadly characterized by impaired ability to interact socially, verbal and nonverbal communication deficits, and a restricted range of behaviors and interests, and it is estimated to affect 6 in 1000 births, with four times as many males diagnosed as females [204, 205]. Autism (or Autism Spectrum Disorders (ASD)) can be divided into idiopathic or familial cases. Familial cases, caused by genetic anomalies including Fragile X and Rett syndromes, tuberous sclerosis, and unbalanced chromosomal rearrangements, make up 10-15% of ASD patients [206, 207]. The remaining 85-90% of cases likely arise from unknown factors that may include diverse CNVs and accompanying gene dosage alterations (chromosomal microdeletions and duplications are reported for 7.4% of ASD cases [206]), helping to account for the heterogeneous phenotypes of autism and ASD [208]. Several cytogenetic abnormalities have been identified in children with ASD by both genome wide association studies and comparative genomic hybridization [208-211] (Table 1). The most frequent recurrent chromosomal abnormality associated with ASD is a duplication of chromosome region 15q11-13, which is found in 1-3% of ASD cases [212]. This abnormality occurs in the same chromosomal region as two other syndromes, Prader-Willi and Angelman, which are paternal or maternal deletions (respectively) leading to distinct phenotypes, including neurodevelopmental deficits and poor motor control, through genomic imprinting [213]. Potential genes in the ASD region of chromosome 15q11-13 are GABRB3 and GABRA5, ɣ -amino butyric acid (GABA) receptor-type A subunits 3 and 5, key receptor subunits in the development of GABAergic interneurons [212, 213]. De novo CNVs associated with autism are rare, but continue to be found and characterized across several chromosomes [209, 214]: 1q21.1 [215]; 2p15-16.1 [216]; 7q22 [217]; 11p12-p13 and 11p14 [214, 218]; 16p11.2 [219, 220]; 17q12 [221]; and 22q11.2 [222]. Regions identified on chromosomes 15, 16, and 22 have been previously mentioned for their association with schizophrenia [191, 194-198], an intriguing connection between two neuropsychiatric disorders that remains to be understood.
The X chromosome has also been extensively studied for connections with ASD. Secondary autism is common in Fragile X patients [223], and men with Klinefelter syndrome (a karyotype of 47, XXY) display more autistic traits compared to non-affected controls [7, 224]. Additionally, low-level mosaic chromosome X gains have been observed by FISH in peripheral blood cells from idiopathic ASD patients [225]. Given the X chromosome disorder links and the observed 4:1 ratio of autism in males versus females, Zhao and colleagues suggest a “unified genetic theory of autism” – two distinct modes of acquiring autism, wherein the male offspring of some families have a 50% risk of ASD because of X-linked dominant transmission with high penetrance, while most male offspring have low risk, and de novo CNVs may contribute to autism [208, 226]. To examine this model, a recent study of families with autistic children from the Autism Genetic Resource Exchange [227] provided evidence that the chromosomal region Xp22.11-p21.2 is linked with an autism predisposition, but inheritance through dominant transmission remains unclear [228]. Proposed gene candidates in this region include the interleukin 1 receptor accessory protein-like gene (IL1RAPL1), which is thought to be involved in neurotransmitter release and synaptic plasticity [228]; other ASD candidates on chromosome X are two neuroligin genes that aid in synapse formation, NLGN3 and NLGN4X, both of which have been associated with autistic behavior [229]. Even with implicated chromosomal regions and potential gene candidates, the etiology of ASD largely remains a mystery. The potential for somatic genomic mosaicism within the brain to account for the varied signals seen in ASD deserves further investigation.
2.5 Mosaic Variegated Aneuploidy (MVA)
MVA is an autosomal recessive disease that, curiously, causes mosaic monosomy and trisomy throughout several tissue types including the brain [230]. While only 35 cases of MVA have been reported worldwide [230-233], clinical manifestations of widespread organismal mosaic aneuploidy are well described, including microcephaly, severe growth and mental retardation, a high risk of cancer, and CNS malformations (e.g., Dandy-Walker complex or migrational defects) [14, 231, 234-236]. Mosaic aneuploidy in MVA arises through premature centromere division – split centromeres and splayed chromatids during metaphase – from mitotic spindle defects, indicated by a lack of metaphase arrest when treated with colcemid [15]. Screening of several MVA pedigrees revealed mutations in BUB1B which encodes a key mitotic spindle checkpoint protein BUBR1 that delays anaphase [15, 230, 233] (Table 1). Wild type BUBR1 inhibits the anaphase-promoting complex/cyclosome (APC/C) checkpoint, preventing cell cycle progression when chromosomes are not properly aligned; mutations have been identified in the kinase domain that phosphorylates APC/C as well as in the domain responsible for recruiting other cell cycle checkpoint proteins [230]. However, BUB1B mutations are not homogenous across all MVA patients, suggesting that other mitotic spindle genes may lead to the same mosaically aneuploid karyotype [11, 230, 232].
Conclusions
Genomic mosaicism within the CNS represents a relatively new frontier towards understanding the development and function of the brain, as well as numerous pathological processes that afflict it. Certain aneuploidies, including extreme forms like MVA and DS, are well recognized for influencing brain function, with clearly demonstrated unambiguous consequences to altering genomic content; studies of mosaic aneuploidy and its consequences in other disease states is only just beginning. Changes observed in gene expression associated with specific aneusomies within a single cell type of normal brain cells implicate functional consequences for aneuploidy in the non-diseased CNS as well, and recent analyses of the developing brain support distinct functions based upon karyotype, with varied aneuploid forms differentially promoting cell survival or death. Studies of aneuploidy in the non-diseased CNS question the assumption that aneuploidy is in fact “abnormal” in the development and function of certain cell lineages, and that it is deleterious – views contradicted by the maintenance of aneuploid populations in the normal brain. Indeed, some forms of aneuploidy may have beneficial consequences for neural development and function; this intriguing hypothesis will be addressed in the future by analyses of living, aneuploid cells. Long-lived cells like post-mitotic neurons may be especially apt at utilizing genomic alterations to their advantage, since they would not be under genomic constraints of highly mitotic cell populations. The stable and seemingly permanent changes produced by genomic alterations in a single neuron could provide a mechanism for creating and stabilizing functional mosaic populations within the brain, such as those constituting a neural network.
Aneuploidies represent a major alteration to neural genomes, but they are certainly not unique in this. The term DNA content variation or “DCV” has been proposed to encompass all of the different forms of genomic changes that are likely to be present within cells of the brain, from aneuploidy to putative mobile LINE elements, de novo CNVs, and other forms of DCV as have been identified within the frontal cortex. DCV within the frontal cortex that is distinct from the pattern observed in the cerebellum from the same individual also demonstrates regional differences in mosaicism, and supports non-random mechanisms in the generation and/or maintenance of this variability. It is notable that types of genomic changes are not mutually exclusive: the genomic landscape could well be heterogeneous with DNA gains, losses, or coincident gains and losses from each of the known sources of DCV.
Our current understanding of the brain does not broadly integrate the existence of genomic mosaicism, but future studies of genomic mosaic alterations using DNA sequences from new single-cell technologies, and strategies that seek to define the cell types, intercellular relationships and global patterns of mosaicism within the brain would complement and expand our knowledge of CNS form and function. Similarly, the possible detection of alterations to normally occurring genomic variation in a disease could identify risk factors, biomarkers and/or new therapeutic targets for the treatment of neurological and psychiatric disorders, particularly for common disease forms that share etiology but not causitive gene associations with rarer familial disease forms. Genomic mosaicism places organizational uniqueness upon the brain, even within syngenic organisms, which could provide a basis for behavioral diversity within a population towards promoting its survival and optimal fitness.
Highlights.
Developing and mature brains are composed of cells that can be genomically varied and mosaic.
Mosaic aneuploidy, as part of DNA content variation (DCV), along with their analyses in brain cells are described.
Functions for neural genomic mosaicism are discussed; roles in cell death and survival implicate selection mechanisms.
The relevance of mosaic genomic variation to neurological and psychiatric disorders is discussed.
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
References
- [1].Alberman E, Mutton D, Morris JK. Cytological and epidemiological findings in trisomies 13, 18, and 21: England and Wales 2004-2009. Am J Med Genet A. 2012;158A:1145–50. doi: 10.1002/ajmg.a.35337. [DOI] [PubMed] [Google Scholar]
- [2].Kalousek DK, Howard-Peebles PN, Olson SB, Barrett IJ, Dorfmann A, Black SH, et al. Confirmation of CVS mosaicism in term placentae and high frequency of intrauterine growth retardation association with confined placental mosaicism. Prenatal Diagn. 1991;11:743–50. doi: 10.1002/pd.1970111002. [DOI] [PubMed] [Google Scholar]
- [3].Benn P, Hsu LY, Perlis T, Schonhaut A. Prenatal diagnosis of chromosome mosaicism. Prenatal Diagn. 1984;4:1–9. doi: 10.1002/pd.1970040102. [DOI] [PubMed] [Google Scholar]
- [4].Jinawath N, Zambrano R, Wohler E, Palmquist MK, Hoover-Fong J, Hamosh A, et al. Mosaic trisomy 13: understanding origin using SNP array. J Med Genet. 2011;48:323–6. doi: 10.1136/jmg.2010.083931. [DOI] [PubMed] [Google Scholar]
- [5].Robinson WP, Binkert F, Bernasconi F, Lorda-Sanchez I, Werder EA, Schinzel AA. Molecular studies of chromosomal mosaicism: relative frequency of chromosome gain or loss and possible role of cell selection. Am J Hum Genet. 1995;56:444–51. [PMC free article] [PubMed] [Google Scholar]
- [6].Cohen FL, Durham JD. Sex chromosome variations in school-age children. J Sch Health. 1985;55:99–102. doi: 10.1111/j.1746-1561.1985.tb04089.x. [DOI] [PubMed] [Google Scholar]
- [7].Visootsak J, Graham JM., Jr. Social function in multiple X and Y chromosome disorders: XXY, XYY, XXYY, XXXY. Dev Disabil Res Rev. 2009;15:328–32. doi: 10.1002/ddrr.76. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [8].Abramsky L, Chapple J. 47,XXY (Klinefelter syndrome) and 47,XYY: estimated rates of and indication for postnatal diagnosis with implications for prenatal counselling. Prenatal Diagn. 1997;17:363–8. doi: 10.1002/(sici)1097-0223(199704)17:4<363::aid-pd79>3.0.co;2-o. [DOI] [PubMed] [Google Scholar]
- [9].Walter E, Mazaika PK, Reiss AL. Insights into brain development from neurogenetic syndromes: evidence from fragile X syndrome, Williams syndrome, Turner syndrome and velocardiofacial syndrome. Neuroscience. 2009;164:257–71. doi: 10.1016/j.neuroscience.2009.04.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [10].Bishop DV, Canning E, Elgar K, Morris E, Jacobs PA, Skuse DH. Distinctive patterns of memory function in subgroups of females with Turner syndrome: evidence for imprinted loci on the X-chromosome affecting neurodevelopment. Neuropsychologia. 2000;38:712–21. doi: 10.1016/s0028-3932(99)00118-9. [DOI] [PubMed] [Google Scholar]
- [11].Garcia-Castillo H, Vasquez-Velasquez AI, Rivera H, Barros-Nunez P. Clinical and genetic heterogeneity in patients with mosaic variegated aneuploidy: delineation of clinical subtypes. Am J Med Genet A. 2008;146A:1687–95. doi: 10.1002/ajmg.a.32315. [DOI] [PubMed] [Google Scholar]
- [12].Kajii T, Kawai T, Takumi T, Misu H, Mabuchi O, Takahashi Y, et al. Mosaic variegated aneuploidy with multiple congenital abnormalities: homozygosity for total premature chromatid separation trait. Am J Med Genet. 1998;78:245–9. doi: 10.1002/(sici)1096-8628(19980707)78:3<245::aid-ajmg7>3.0.co;2-o. [DOI] [PubMed] [Google Scholar]
- [13].Warburton D, Anyane-Yeboa K, Taterka P, Yu CY, Olsen D. Mosaic variegated aneuploidy with microcephaly: a new human mitotic mutant? Ann Genet. 1991;34:287–92. [PubMed] [Google Scholar]
- [14].Kajii T, Ikeuchi T, Yang ZQ, Nakamura Y, Tsuji Y, Yokomori K, et al. Cancer-prone syndrome of mosaic variegated aneuploidy and total premature chromatid separation: report of five infants. Am J Med Genet. 2001;104:57–64. doi: 10.1002/ajmg.1580. [DOI] [PubMed] [Google Scholar]
- [15].Matsuura S, Ito E, Tauchi H, Komatsu K, Ikeuchi T, Kajii T. Chromosomal instability syndrome of total premature chromatid separation with mosaic variegated aneuploidy is defective in mitotic-spindle checkpoint. Am J Hum Genet. 2000;67:483–6. doi: 10.1086/303022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [16].Iourov IY, Vorsanova SG, Liehr T, Yurov YB. Aneuploidy in the normal, Alzheimer’s disease and ataxia-telangiectasia brain: differential expression and pathological meaning. Neurobiol Dis. 2009;34:212–20. doi: 10.1016/j.nbd.2009.01.003. [DOI] [PubMed] [Google Scholar]
- [17].Kaushal D, Contos JJ, Treuner K, Yang AH, Kingsbury MA, Rehen SK, et al. Alteration of gene expression by chromosome loss in the postnatal mouse brain. J Neurosci. 2003;23:5599–606. doi: 10.1523/JNEUROSCI.23-13-05599.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [18].Kingsbury MA, Friedman B, McConnell MJ, Rehen SK, Yang AH, Kaushal D, et al. Aneuploid neurons are functionally active and integrated into brain circuitry. Proc Natl Acad Sci U S A. 2005;102:6143–7. doi: 10.1073/pnas.0408171102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [19].Kingsbury MA, Yung YC, Peterson SE, Westra JW, Chun J. Aneuploidy in the normal and diseased brain. Cell Mol Life Sci. 2006;63:2626–41. doi: 10.1007/s00018-006-6169-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [20].Pack SD, Weil RJ, Vortmeyer AO, Zeng W, Li J, Okamoto H, et al. Individual adult human neurons display aneuploidy: detection by fluorescence in situ hybridization and single neuron PCR. Cell Cycle. 2005;4:1758–60. doi: 10.4161/cc.4.12.2153. [DOI] [PubMed] [Google Scholar]
- [21].Rajendran RS, Wellbrock UM, Zupanc GK. Apoptotic cell death, long-term persistence, and neuronal differentiation of aneuploid cells generated in the adult brain of teleost fish. Dev Neurobiol. 2008;68:1257–68. doi: 10.1002/dneu.20656. [DOI] [PubMed] [Google Scholar]
- [22].Rajendran RS, Zupanc MM, Losche A, Westra J, Chun J, Zupanc GK. Numerical chromosome variation and mitotic segregation defects in the adult brain of teleost fish. Dev Neurobiol. 2007;67:1334–47. doi: 10.1002/dneu.20365. [DOI] [PubMed] [Google Scholar]
- [23].Rehen SK, McConnell MJ, Kaushal D, Kingsbury MA, Yang AH, Chun J. Chromosomal variation in neurons of the developing and adult mammalian nervous system. Proc Natl Acad Sci U S A. 2001;98:13361–6. doi: 10.1073/pnas.231487398. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [24].Rehen SK, Yung YC, McCreight MP, Kaushal D, Yang AH, Almeida BS, et al. Constitutional aneuploidy in the normal human brain. J Neurosci. 2005;25:2176–80. doi: 10.1523/JNEUROSCI.4560-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [25].Westra JW, Peterson SE, Yung YC, Mutoh T, Barral S, Chun J. Aneuploid mosaicism in the developing and adult cerebellar cortex. J Comp Neurol. 2008;507:1944–51. doi: 10.1002/cne.21648. [DOI] [PubMed] [Google Scholar]
- [26].Yang AH, Kaushal D, Rehen SK, Kriedt K, Kingsbury MA, McConnell MJ, et al. Chromosome segregation defects contribute to aneuploidy in normal neural progenitor cells. J Neurosci. 2003;23:10454–62. doi: 10.1523/JNEUROSCI.23-32-10454.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [27].Yurov YB, Iourov IY, Monakhov VV, Soloviev IV, Vostrikov VM, Vorsanova SG. The variation of aneuploidy frequency in the developing and adult human brain revealed by an interphase FISH study. J Histochem Cytochem. 2005;53:385–90. doi: 10.1369/jhc.4A6430.2005. [DOI] [PubMed] [Google Scholar]
- [28].Yurov YB, Iourov IY, Vorsanova SG, Liehr T, Kolotii AD, Kutsev SI, et al. Aneuploidy and confined chromosomal mosaicism in the developing human brain. PLoS ONE. 2007;2:e558. doi: 10.1371/journal.pone.0000558. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [29].Westra JW, Rivera RR, Bushman DM, Yung YC, Peterson SE, Barral S, et al. Neuronal DNA content variation (DCV) with regional and individual differences in the human brain. J Comp Neurol. 2010;518:3981–4000. doi: 10.1002/cne.22436. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [30].Boveri T. Concerning the origin of malignant tumours by Theodor Boveri. Translated and annotated by Henry Harris. J Cell Sci. 2008;121(Suppl 1):1–84. doi: 10.1242/jcs.025742. [DOI] [PubMed] [Google Scholar]
- [31].Barch MJ, Knutsen T, Spurbeck JL. The AGT Cytogenetics Laboratory Manual. Lippincott; Philadelphia: 1997. [Google Scholar]
- [32].Tjio JH, Levan A. THE CHROMOSOME NUMBER OF MAN. Hereditas. 1956;42:1–6. [Google Scholar]
- [33].Watson JD, Crick FH. Molecular structure of nucleic acids; a structure for deoxyribose nucleic acid. Nature. 1953;171:737–8. doi: 10.1038/171737a0. [DOI] [PubMed] [Google Scholar]
- [34].Schrock E, du Manoir S, Veldman T, Schoell B, Wienberg J, Ferguson-Smith MA, et al. Multicolor spectral karyotyping of human chromosomes. Science. 1996;273:494–7. doi: 10.1126/science.273.5274.494. [DOI] [PubMed] [Google Scholar]
- [35].Iourov IY, Soloviev IV, Vorsanova SG, Monakhov VV, Yurov YB. An approach for quantitative assessment of fluorescence in situ hybridization (FISH) signals for applied human molecular cytogenetics. J Histochem Cytochem. 2005;53:401–8. doi: 10.1369/jhc.4A6419.2005. [DOI] [PubMed] [Google Scholar]
- [36].Liehr T, Heller A, Starke H, Rubtsov N, Trifonov V, Mrasek K, et al. Microdissection based high resolution multicolor banding for all 24 human chromosomes. Int J Mol Med. 2002;9:335–9. [PubMed] [Google Scholar]
- [37].Iourov IY, Vorsanova SG, Yurov YB. Chromosomal variation in mammalian neuronal cells: known facts and attractive hypotheses. Int Rev Cytol. 2006;249:143–91. doi: 10.1016/S0074-7696(06)49003-3. [DOI] [PubMed] [Google Scholar]
- [38].Gorman P, Roylance R. Fluorescence in situ hybridization and comparative genomic hybridization. Methods Mol Med. 2006;120:269–95. doi: 10.1385/1-59259-969-9:269. [DOI] [PubMed] [Google Scholar]
- [39].Pinkel D, Albertson DG. Comparative genomic hybridization. Annu Rev Genomics Hum Genet. 2005;6:331–54. doi: 10.1146/annurev.genom.6.080604.162140. [DOI] [PubMed] [Google Scholar]
- [40].Ballif BC, Rorem EA, Sundin K, Lincicum M, Gaskin S, Coppinger J, et al. Detection of low-level mosaicism by array CGH in routine diagnostic specimens. Am J Med Genet A. 2006;140:2757–67. doi: 10.1002/ajmg.a.31539. [DOI] [PubMed] [Google Scholar]
- [41].Spits C, Le Caignec C, De Rycke M, Van Haute L, Van Steirteghem A, Liebaers I, et al. Whole-genome multiple displacement amplification from single cells. Nat Protoc. 2006;1:1965–70. doi: 10.1038/nprot.2006.326. [DOI] [PubMed] [Google Scholar]
- [42].Navin N, Hicks J. Future medical applications of single-cell sequencing in cancer. Genome Med. 2011;3:31. doi: 10.1186/gm247. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [43].Navin N, Kendall J, Troge J, Andrews P, Rodgers L, McIndoo J, et al. Tumour evolution inferred by single-cell sequencing. Nature. 2011;472:90–4. doi: 10.1038/nature09807. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [44].Evrony GD, Cai X, Lee E, Hills LB, Elhosary PC, Lehmann HS, et al. Single-neuron sequencing analysis of l1 retrotransposition and somatic mutation in the human brain. Cell. 2012;151:483–96. doi: 10.1016/j.cell.2012.09.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [45].Johnson LA. Sex preselection by flow cytometric separation of X and Y chromosome-bearing sperm based on DNA difference: a review. Reprod Fertil Dev. 1995;7:893–903. doi: 10.1071/rd9950893. [DOI] [PubMed] [Google Scholar]
- [46].Laerum OD, Farsund T. Clinical application of flow cytometry: a review. Cytometry. 1981;2:1–13. doi: 10.1002/cyto.990020102. [DOI] [PubMed] [Google Scholar]
- [47].Peterson SE, Chun J, Loring J. FISH analysis of human pluripotent stem cells. Methods Mol Biol. 2011;767:191–200. doi: 10.1007/978-1-61779-201-4_14. [DOI] [PubMed] [Google Scholar]
- [48].Peterson SE, Westra JW, Rehen SK, Young H, Bushman DM, Paczkowski CM, et al. Normal human pluripotent stem cell lines exhibit pervasive mosaic aneuploidy. PLoS ONE. 2011;6:e23018. doi: 10.1371/journal.pone.0023018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [49].Hassold T, Hunt P. To err (meiotically) is human: the genesis of human aneuploidy. Nat Rev Genet. 2001;2:280–91. doi: 10.1038/35066065. [DOI] [PubMed] [Google Scholar]
- [50].Hughes TR, Roberts CJ, Dai H, Jones AR, Meyer MR, Slade D, et al. Widespread aneuploidy revealed by DNA microarray expression profiling. Nat Genet. 2000;25:333–7. doi: 10.1038/77116. [DOI] [PubMed] [Google Scholar]
- [51].Sheltzer JM, Torres EM, Dunham MJ, Amon A. Transcriptional consequences of aneuploidy. Proc Natl Acad Sci U S A. 2012;109:12644–9. doi: 10.1073/pnas.1209227109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [52].Torres EM, Williams BR, Amon A. Aneuploidy: cells losing their balance. Genetics. 2008;179:737–46. doi: 10.1534/genetics.108.090878. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [53].Yurov YB, Iourov IY, Monakhov VV, Soloviev IV, Vostrikov VM, Vorsanova SG. The variation of aneuploidy frequency in the developing and adult human brain revealed by an interphase FISH study. J Histochem Cytochem. 2005;53:385–90. doi: 10.1369/jhc.4A6430.2005. [DOI] [PubMed] [Google Scholar]
- [54].Iourov IY, Liehr T, Vorsanova SG, Kolotii AD, Yurov YB. Visualization of interphase chromosomes in postmitotic cells of the human brain by multicolour banding (MCB) Chromosome Res. 2006;14:223–9. doi: 10.1007/s10577-006-1037-6. [DOI] [PubMed] [Google Scholar]
- [55].Lengauer C, Kinzler KW, Vogelstein B. Genetic instabilities in human cancers. Nature. 1998;396:643–9. doi: 10.1038/25292. [DOI] [PubMed] [Google Scholar]
- [56].Antonarakis SE, Lyle R, Dermitzakis ET, Reymond A, Deutsch S. Chromosome 21 and down syndrome: from genomics to pathophysiology. Nat Rev Genet. 2004;5:725–38. doi: 10.1038/nrg1448. [DOI] [PubMed] [Google Scholar]
- [57].Duesberg P, Rasnick D. Aneuploidy, the somatic mutation that makes cancer a species of its own. Cell Motil Cytoskeleton. 2000;47:81–107. doi: 10.1002/1097-0169(200010)47:2<81::AID-CM1>3.0.CO;2-#. [DOI] [PubMed] [Google Scholar]
- [58].Torres EM, Sokolsky T, Tucker CM, Chan LY, Boselli M, Dunham MJ, et al. Effects of aneuploidy on cellular physiology and cell division in haploid yeast. Science. 2007;317:916–24. doi: 10.1126/science.1142210. [DOI] [PubMed] [Google Scholar]
- [59].Sheltzer JM, Blank HM, Pfau SJ, Tange Y, George BM, Humpton TJ, et al. Aneuploidy drives genomic instability in yeast. Science. 2011;333:1026–30. doi: 10.1126/science.1206412. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [60].Humpherys D, Eggan K, Akutsu H, Hochedlinger K, Rideout WM, 3rd, Biniszkiewicz D, et al. Epigenetic instability in ES cells and cloned mice. Science. 2001;293:95–7. doi: 10.1126/science.1061402. [DOI] [PubMed] [Google Scholar]
- [61].Williams BR, Prabhu VR, Hunter KE, Glazier CM, Whittaker CA, Housman DE, et al. Aneuploidy affects proliferation and spontaneous immortalization in mammalian cells. Science. 2008;322:703–9. doi: 10.1126/science.1160058. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [62].Pavelka N, Rancati G, Zhu J, Bradford WD, Saraf A, Florens L, et al. Aneuploidy confers quantitative proteome changes and phenotypic variation in budding yeast. Nature. 2010;468:321–5. doi: 10.1038/nature09529. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [63].Torres EM, Dephoure N, Panneerselvam A, Tucker CM, Whittaker CA, Gygi SP, et al. Identification of aneuploidy-tolerating mutations. Cell. 2010;143:71–83. doi: 10.1016/j.cell.2010.08.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [64].Sheltzer JM, Amon A. The aneuploidy paradox: costs and benefits of an incorrect karyotype. Trends Genet. 2011;27:446–53. doi: 10.1016/j.tig.2011.07.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [65].Peterson SE, Yang AH, Bushman DM, Westra JW, Yung YC, Barral S, et al. Aneuploid cells are differentially susceptible to caspase-mediated death during embryonic cerebral cortical development. J Neurosci. 2012;32:16213–22. doi: 10.1523/JNEUROSCI.3706-12.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [66].Chun J. Selected comparison of immune and nervous system development. Adv Immunol. 2001;77:297–322. doi: 10.1016/s0065-2776(01)77020-4. [DOI] [PubMed] [Google Scholar]
- [67].Haydar TF, Kuan CY, Flavell RA, Rakic P. The role of cell death in regulating the size and shape of the mammalian forebrain. Cereb Cortex. 1999;9:621–6. doi: 10.1093/cercor/9.6.621. [DOI] [PubMed] [Google Scholar]
- [68].Kuan CY, Roth KA, Flavell RA, Rakic P. Mechanisms of programmed cell death in the developing brain. Trends Neurosci. 2000;23:291–7. doi: 10.1016/s0166-2236(00)01581-2. [DOI] [PubMed] [Google Scholar]
- [69].Rakic P. Less is more: progenitor death and cortical size. Nat Neurosci. 2005;8:981–2. doi: 10.1038/nn0805-981. [DOI] [PubMed] [Google Scholar]
- [70].McConnell MJ, Macmillan HR, Chun J. Mathematical modeling supports substantial mouse neural progenitor cell death. Neural Dev. 2009;4:28. doi: 10.1186/1749-8104-4-28. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [71].Yung YC, Kennedy G, Chun J. Identification of neural programmed cell death through the detection of DNA fragmentation in situ and by PCR. Curr Protoc Neurosci. 2009 doi: 10.1002/0471142301.ns0308s48. Chapter 3:Unit 3 8. [DOI] [PubMed] [Google Scholar]
- [72].Kuida K, Haydar TF, Kuan CY, Gu Y, Taya C, Karasuyama H, et al. Reduced apoptosis and cytochrome c-mediated caspase activation in mice lacking caspase 9. Cell. 1998;94:325–37. doi: 10.1016/s0092-8674(00)81476-2. [DOI] [PubMed] [Google Scholar]
- [73].Kuida K, Zheng TS, Na S, Kuan C, Yang D, Karasuyama H, et al. Decreased apoptosis in the brain and premature lethality in CPP32-deficient mice. Nature. 1996;384:368–72. doi: 10.1038/384368a0. [DOI] [PubMed] [Google Scholar]
- [74].Hakem R, Hakem A, Duncan GS, Henderson JT, Woo M, Soengas MS, et al. Differential requirement for caspase 9 in apoptotic pathways in vivo. Cell. 1998;94:339–52. doi: 10.1016/s0092-8674(00)81477-4. [DOI] [PubMed] [Google Scholar]
- [75].Pompeiano M, Blaschke AJ, Flavell RA, Srinivasan A, Chun J. Decreased apoptosis in proliferative and postmitotic regions of the Caspase 3-deficient embryonic central nervous system. J Comp Neurol. 2000;423:1–12. [PubMed] [Google Scholar]
- [76].Oppenheim RW, Flavell RA, Vinsant S, Prevette D, Kuan CY, Rakic P. Programmed cell death of developing mammalian neurons after genetic deletion of caspases. J Neurosci. 2001;21:4752–60. doi: 10.1523/JNEUROSCI.21-13-04752.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [77].Leonard JR, Klocke BJ, D’Sa C, Flavell RA, Roth KA. Strain-dependent neurodevelopmental abnormalities in caspase-3-deficient mice. J Neuropathol Exp Neurol. 2002;61:673–7. doi: 10.1093/jnen/61.8.673. [DOI] [PubMed] [Google Scholar]
- [78].Momoi T, Fujita E, Urase K. Strain-specific caspase-3-dependent programmed cell death in the early developing mouse forebrain. Neuroreport. 2003;14:111–5. doi: 10.1097/00001756-200301200-00021. [DOI] [PubMed] [Google Scholar]
- [79].Devalle S, Sartore RC, Paulsen BS, Borges HL, Martins RA, Rehen SK. Implications of aneuploidy for stem cell biology and brain therapeutics. Front Cell Neurosci. 2012;6:36. doi: 10.3389/fncel.2012.00036. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [80].Muotri AR, Chu VT, Marchetto MC, Deng W, Moran JV, Gage FH. Somatic mosaicism in neuronal precursor cells mediated by L1 retrotransposition. Nature. 2005;435:903–10. doi: 10.1038/nature03663. [DOI] [PubMed] [Google Scholar]
- [81].Muotri AR, Gage FH. Generation of neuronal variability and complexity. Nature. 2006;441:1087–93. doi: 10.1038/nature04959. [DOI] [PubMed] [Google Scholar]
- [82].Sebat J, Lakshmi B, Troge J, Alexander J, Young J, Lundin P, et al. Large-scale copy number polymorphism in the human genome. Science. 2004;305:525–8. doi: 10.1126/science.1098918. [DOI] [PubMed] [Google Scholar]
- [83].Hussein SM, Batada NN, Vuoristo S, Ching RW, Autio R, Narva E, et al. Copy number variation and selection during reprogramming to pluripotency. Nature. 2011;471:58–62. doi: 10.1038/nature09871. [DOI] [PubMed] [Google Scholar]
- [84].Gore A, Li Z, Fung HL, Young JE, Agarwal S, Antosiewicz-Bourget J, et al. Somatic coding mutations in human induced pluripotent stem cells. Nature. 2011;471:63–7. doi: 10.1038/nature09805. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [85].Longo L, Bygrave A, Grosveld FG, Pandolfi PP. The chromosome make-up of mouse embryonic stem cells is predictive of somatic and germ cell chimaerism. Transgenic Res. 1997;6:321–8. doi: 10.1023/a:1018418914106. [DOI] [PubMed] [Google Scholar]
- [86].Cervantes RB, Stringer JR, Shao C, Tischfield JA, Stambrook PJ. Embryonic stem cells and somatic cells differ in mutation frequency and type. Proc Natl Acad Sci U S A. 2002;99:3586–90. doi: 10.1073/pnas.062527199. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [87].Eggan K, Rode A, Jentsch I, Samuel C, Hennek T, Tintrup H, et al. Male and female mice derived from the same embryonic stem cell clone by tetraploid embryo complementation. Nat Biotechnol. 2002;20:455–9. doi: 10.1038/nbt0502-455. [DOI] [PubMed] [Google Scholar]
- [88].Draper JS, Smith K, Gokhale P, Moore HD, Maltby E, Johnson J, et al. Recurrent gain of chromosomes 17q and 12 in cultured human embryonic stem cells. Nat Biotechnol. 2004;22:53–4. doi: 10.1038/nbt922. [DOI] [PubMed] [Google Scholar]
- [89].Buzzard JJ, Gough NM, Crook JM, Colman A. Karyotype of human ES cells during extended culture. Nat Biotechnol. 2004;22:381–2. doi: 10.1038/nbt0404-381. author reply 2. [DOI] [PubMed] [Google Scholar]
- [90].Chang HH, Hemberg M, Barahona M, Ingber DE, Huang S. Transcriptome-wide noise controls lineage choice in mammalian progenitor cells. Nature. 2008;453:544–7. doi: 10.1038/nature06965. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [91].Graf T, Stadtfeld M. Heterogeneity of embryonic and adult stem cells. Cell Stem Cell. 2008;3:480–3. doi: 10.1016/j.stem.2008.10.007. [DOI] [PubMed] [Google Scholar]
- [92].Lengauer C, Kinzler KW, Vogelstein B. Genetic instability in colorectal cancers. Nature. 1997;386:623–7. doi: 10.1038/386623a0. [DOI] [PubMed] [Google Scholar]
- [93].Rajagopalan H, Lengauer C. Aneuploidy and cancer. Nature. 2004;432:338–41. doi: 10.1038/nature03099. [DOI] [PubMed] [Google Scholar]
- [94].Takeuchi H, Kubota T, Kitai R, Matsuda K, Hashimoto N, Sato K. Chromosome 1p and 19q deletions in malignant glioneuronal tumors with oligodendroglioma-like component. J Neurooncol. 2009;91:33–8. doi: 10.1007/s11060-008-9690-6. [DOI] [PubMed] [Google Scholar]
- [95].Thiel G, Losanowa T, Kintzel D, Nisch G, Martin H, Vorpahl K, et al. Karyotypes in 90 human gliomas. Cancer Genet Cytogenet. 1992;58:109–20. doi: 10.1016/0165-4608(92)90095-p. [DOI] [PubMed] [Google Scholar]
- [96].Baker DE, Harrison NJ, Maltby E, Smith K, Moore HD, Shaw PJ, et al. Adaptation to culture of human embryonic stem cells and oncogenesis in vivo. Nat Biotechnol. 2007;25:207–15. doi: 10.1038/nbt1285. [DOI] [PubMed] [Google Scholar]
- [97].Hovatta O, Jaconi M, Tohonen V, Bena F, Gimelli S, Bosman A, et al. A teratocarcinoma-like human embryonic stem cell (hESC) line and four hESC lines reveal potentially oncogenic genomic changes. PLoS ONE. 2010;5:e10263. doi: 10.1371/journal.pone.0010263. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [98].Patau K, Smith DW, Therman E, Inhorn SL, Wagner HP. Multiple congenital anomaly caused by an extra autosome. Lancet. 1960;1:790–3. doi: 10.1016/s0140-6736(60)90676-0. [DOI] [PubMed] [Google Scholar]
- [99].Edwards JH, Harnden DG, Cameron AH, Crosse VM, Wolff OH. A new trisomic syndrome. Lancet. 1960;1:787–90. doi: 10.1016/s0140-6736(60)90675-9. [DOI] [PubMed] [Google Scholar]
- [100].Lejeune J, Gautier M, Turpin R. [Study of somatic chromosomes from 9 mongoloid children] C R Hebd Seances Acad Sci. 1959;248:1721–2. [PubMed] [Google Scholar]
- [101].Lejeune J, Turpin R, Gautier M. [Mongolism; a chromosomal disease (trisomy)] Bull Acad Natl Med. 1959;143:256–65. [PubMed] [Google Scholar]
- [102].Lejeune J, Turpin R, Gautier M. [Chromosomic diagnosis of mongolism] Arch Fr Pediatr. 1959;16:962–3. [PubMed] [Google Scholar]
- [103].Wisniewski KE. Down syndrome children often have brain with maturation delay, retardation of growth, and cortical dysgenesis. Am J Med Genet Suppl. 1990;7:274–81. doi: 10.1002/ajmg.1320370755. [DOI] [PubMed] [Google Scholar]
- [104].Patterson D, Costa AC. Down syndrome and genetics - a case of linked histories. Nat Rev Genet. 2005;6:137–47. doi: 10.1038/nrg1525. [DOI] [PubMed] [Google Scholar]
- [105].Mao R, Zielke CL, Zielke HR, Pevsner J. Global up-regulation of chromosome 21 gene expression in the developing Down syndrome brain. Genomics. 2003;81:457–67. doi: 10.1016/s0888-7543(03)00035-1. [DOI] [PubMed] [Google Scholar]
- [106].Sawa A. Alteration of gene expression in Down’s syndrome (DS) brains: its significance in neurodegeneration. J Neural Transm Suppl. 2001:361–71. doi: 10.1007/978-3-7091-6262-0_30. [DOI] [PubMed] [Google Scholar]
- [107].Park J, Oh Y, Chung KC. Two key genes closely implicated with the neuropathological characteristics in Down syndrome: DYRK1A and RCAN1. BMB Rep. 2009;42:6–15. doi: 10.5483/bmbrep.2009.42.1.006. [DOI] [PubMed] [Google Scholar]
- [108].Lichter P, Cremer T, Tang CJ, Watkins PC, Manuelidis L, Ward DC. Rapid detection of human chromosome 21 aberrations by in situ hybridization. Proc Natl Acad Sci U S A. 1988;85:9664–8. doi: 10.1073/pnas.85.24.9664. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [109].Stoll C, Alembik Y, Dott B, Roth MP. Study of Down syndrome in 238,942 consecutive births. Ann Genet. 1998;41:44–51. [PubMed] [Google Scholar]
- [110].Ermak G, Pritchard MA, Dronjak S, Niu B, Davies KJ. Do RCAN1 proteins link chronic stress with neurodegeneration? FASEB J. 2011;25:3306–11. doi: 10.1096/fj.11-185728. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [111].Porta S, Serra SA, Huch M, Valverde MA, Llorens F, Estivill X, et al. RCAN1 (DSCR1) increases neuronal susceptibility to oxidative stress: a potential pathogenic process in neurodegeneration. Hum Mol Genet. 2007;16:1039–50. doi: 10.1093/hmg/ddm049. [DOI] [PubMed] [Google Scholar]
- [112].Wegiel J, Gong CX, Hwang YW. The role of DYRK1A in neurodegenerative diseases. FEBS J. 2011;278:236–45. doi: 10.1111/j.1742-4658.2010.07955.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [113].Wegiel J, Kaczmarski W, Barua M, Kuchna I, Nowicki K, Wang KC, et al. Link between DYRK1A overexpression and several-fold enhancement of neurofibrillary degeneration with 3-repeat tau protein in Down syndrome. J Neuropathol Exp Neurol. 2011;70:36–50. doi: 10.1097/NEN.0b013e318202bfa1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [114].Rachidi M, Lopes C. Mental retardation in Down syndrome: from gene dosage imbalance to molecular and cellular mechanisms. Neurosci Res. 2007;59:349–69. doi: 10.1016/j.neures.2007.08.007. [DOI] [PubMed] [Google Scholar]
- [115].Michiels C, Raes M, Toussaint O, Remacle J. Importance of Se-glutathione peroxidase, catalase, and Cu/Zn-SOD for cell survival against oxidative stress. Free Radical Biol Med. 1994;17:235–48. doi: 10.1016/0891-5849(94)90079-5. [DOI] [PubMed] [Google Scholar]
- [116].Muchova J, Sustrova M, Garaiova I, Liptakova A, Blazicek P, Kvasnicka P, et al. Influence of age on activities of antioxidant enzymes and lipid peroxidation products in erythrocytes and neutrophils of Down syndrome patients. Free Radical Biol Med. 2001;31:499–508. doi: 10.1016/s0891-5849(01)00609-8. [DOI] [PubMed] [Google Scholar]
- [117].Sulthana SM, Kumar SN, Sridhar MG, Bhat BV, Rao KR. Levels of Non Enzymatic Antioxidants in Down Syndrome. Indian J Pediatr. 2012 doi: 10.1007/s12098-012-0795-8. [DOI] [PubMed] [Google Scholar]
- [118].Burhans WC, Weinberger M. DNA replication stress, genome instability and aging. Nucleic Acids Res. 2007;35:7545–56. doi: 10.1093/nar/gkm1059. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [119].Taupin P. A dual activity of ROS and oxidative stress on adult neurogenesis and Alzheimer’s disease. Cent Nerv Syst Agents Med Chem. 2010;10:16–21. doi: 10.2174/187152410790780172. [DOI] [PubMed] [Google Scholar]
- [120].Ermak G, Morgan TE, Davies KJ. Chronic overexpression of the calcineurin inhibitory gene DSCR1 (Adapt78) is associated with Alzheimer’s disease. J Biol Chem. 2001;276:38787–94. doi: 10.1074/jbc.M102829200. [DOI] [PubMed] [Google Scholar]
- [121].Harris CD, Ermak G, Davies KJ. RCAN1-1L is overexpressed in neurons of Alzheimer’s disease patients. FEBS J. 2007;274:1715–24. doi: 10.1111/j.1742-4658.2007.05717.x. [DOI] [PubMed] [Google Scholar]
- [122].Fuentes JJ, Genesca L, Kingsbury TJ, Cunningham KW, Perez-Riba M, Estivill X, et al. DSCR1, overexpressed in Down syndrome, is an inhibitor of calcineurin-mediated signaling pathways. Hum Mol Genet. 2000;9:1681–90. doi: 10.1093/hmg/9.11.1681. [DOI] [PubMed] [Google Scholar]
- [123].Crawford DR, Leahy KP, Abramova N, Lan L, Wang Y, Davies KJ. Hamster adapt78 mRNA is a Down syndrome critical region homologue that is inducible by oxidative stress. Arch Biochem Biophys. 1997;342:6–12. doi: 10.1006/abbi.1997.0109. [DOI] [PubMed] [Google Scholar]
- [124].Sun X, Wu Y, Chen B, Zhang Z, Zhou W, Tong Y, et al. Regulator of calcineurin 1 (RCAN1) facilitates neuronal apoptosis through caspase-3 activation. J Biol Chem. 2011;286:9049–62. doi: 10.1074/jbc.M110.177519. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [125].Wu Y, Song W. Regulation of RCAN1 translation and its role in oxidative stress-induced apoptosis. FASEB J. 2012 doi: 10.1096/fj.12-213124. [DOI] [PubMed] [Google Scholar]
- [126].Ermak G, Harris CD, Battocchio D, Davies KJ. RCAN1 (DSCR1 or Adapt78) stimulates expression of GSK-3beta. FEBS J. 2006;273:2100–9. doi: 10.1111/j.1742-4658.2006.05217.x. [DOI] [PubMed] [Google Scholar]
- [127].Poppek D, Keck S, Ermak G, Jung T, Stolzing A, Ullrich O, et al. Phosphorylation inhibits turnover of the tau protein by the proteasome: influence of RCAN1 and oxidative stress. Biochem J. 2006;400:511–20. doi: 10.1042/BJ20060463. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [128].Tejedor FJ, Hammerle B. MNB/DYRK1A as a multiple regulator of neuronal development. FEBS J. 2011;278:223–35. doi: 10.1111/j.1742-4658.2010.07954.x. [DOI] [PubMed] [Google Scholar]
- [129].Altafaj X, Dierssen M, Baamonde C, Marti E, Visa J, Guimera J, et al. Neurodevelopmental delay, motor abnormalities and cognitive deficits in transgenic mice overexpressing Dyrk1A (minibrain), a murine model of Down’s syndrome. Hum Mol Genet. 2001;10:1915–23. doi: 10.1093/hmg/10.18.1915. [DOI] [PubMed] [Google Scholar]
- [130].Arron JR, Winslow MM, Polleri A, Chang CP, Wu H, Gao X, et al. 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]
- [131].Liu F, Liang Z, Wegiel J, Hwang YW, Iqbal K, Grundke-Iqbal I, et al. Overexpression of Dyrk1A contributes to neurofibrillary degeneration in Down syndrome. FASEB J. 2008;22:3224–33. doi: 10.1096/fj.07-104539. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [132].Wegiel J, Dowjat K, Kaczmarski W, Kuchna I, Nowicki K, Frackowiak J, et al. The role of overexpressed DYRK1A protein in the early onset of neurofibrillary degeneration in Down syndrome. Acta Neuropathol (Berl) 2008;116:391–407. doi: 10.1007/s00401-008-0419-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [133].Ding S, Shi J, Qian W, Iqbal K, Grundke-Iqbal I, Gong CX, et al. Regulation of alternative splicing of tau exon 10 by 9G8 and Dyrk1A. Neurobiol Aging. 2012;33:1389–99. doi: 10.1016/j.neurobiolaging.2010.11.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [134].Yin X, Jin N, Gu J, Shi J, Zhou J, Gong CX, et al. Dual-specificity tyrosine phosphorylation-regulated kinase 1A (Dyrk1A) modulates serine/arginine-rich protein 55 (SRp55)-promoted Tau exon 10 inclusion. J Biol Chem. 2012;287:30497–506. doi: 10.1074/jbc.M112.355412. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [135].Alzheimer’s Association . 2012 Alzheimer’s Disease Facts and Figures. Alzheimer’s and Dementia; 2012. [DOI] [PubMed] [Google Scholar]
- [136].Gatz M, Reynolds CA, Fratiglioni L, Johansson B, Mortimer JA, Berg S, et al. Role of genes and environments for explaining Alzheimer disease. Arch Gen Psychiatry. 2006;63:168–74. doi: 10.1001/archpsyc.63.2.168. [DOI] [PubMed] [Google Scholar]
- [137].Bertram L, Lill CM, Tanzi RE. The genetics of Alzheimer disease: back to the future. Neuron. 2010;68:270–81. doi: 10.1016/j.neuron.2010.10.013. [DOI] [PubMed] [Google Scholar]
- [138].Herrup K. Reimagining Alzheimer’s disease--an age-based hypothesis. J Neurosci. 2010;30:16755–62. doi: 10.1523/JNEUROSCI.4521-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [139].Hardy J, Selkoe DJ. The amyloid hypothesis of Alzheimer’s disease: progress and problems on the road to therapeutics. Science. 2002;297:353–6. doi: 10.1126/science.1072994. [DOI] [PubMed] [Google Scholar]
- [140].Glenner GG, Wong CW. Alzheimer’s disease and Down’s syndrome: sharing of a unique cerebrovascular amyloid fibril protein. Biochem Biophys Res Commun. 1984;122:1131–5. doi: 10.1016/0006-291x(84)91209-9. [DOI] [PubMed] [Google Scholar]
- [141].Masters CL, Multhaup G, Simms G, Pottgiesser J, Martins RN, Beyreuther K. Neuronal origin of a cerebral amyloid: neurofibrillary tangles of Alzheimer’s disease contain the same protein as the amyloid of plaque cores and blood vessels. EMBO J. 1985;4:2757–63. doi: 10.1002/j.1460-2075.1985.tb04000.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [142].Goldgaber D, Lerman MI, McBride OW, Saffiotti U, Gajdusek DC. Characterization and chromosomal localization of a cDNA encoding brain amyloid of Alzheimer’s disease. Science. 1987;235:877–80. doi: 10.1126/science.3810169. [DOI] [PubMed] [Google Scholar]
- [143].St George-Hyslop PH, Tanzi RE, Polinsky RJ, Haines JL, Nee L, Watkins PC, et al. The genetic defect causing familial Alzheimer’s disease maps on chromosome 21. Science. 1987;235:885–90. doi: 10.1126/science.2880399. [DOI] [PubMed] [Google Scholar]
- [144].Tanzi RE, Gusella JF, Watkins PC, Bruns GA, St George-Hyslop P, Van Keuren ML, et al. Amyloid beta protein gene: cDNA, mRNA distribution, and genetic linkage near the Alzheimer locus. Science. 1987;235:880–4. doi: 10.1126/science.2949367. [DOI] [PubMed] [Google Scholar]
- [145].Potter H. Review and hypothesis: Alzheimer disease and Down syndrome--chromosome 21 nondisjunction may underlie both disorders. Am J Hum Genet. 1991;48:1192–200. [PMC free article] [PubMed] [Google Scholar]
- [146].Tanzi RE, Bird ED, Latt SA, Neve RL. The amyloid beta protein gene is not duplicated in brains from patients with Alzheimer’s disease. Science. 1987;238:666–9. doi: 10.1126/science.2890207. [DOI] [PubMed] [Google Scholar]
- [147].St George-Hyslop PH, Tanzi RE, Polinsky RJ, Neve RL, Pollen D, Drachman D, et al. Absence of duplication of chromosome 21 genes in familial and sporadic Alzheimer’s disease. Science. 1987;238:664–6. doi: 10.1126/science.2890206. [DOI] [PubMed] [Google Scholar]
- [148].Rovelet-Lecrux A, Hannequin D, Raux G, Le Meur N, Laquerriere A, Vital A, et al. APP locus duplication causes autosomal dominant early-onszheimer disease with cerebral amyloid angiopathy. Nat Genet. 2006;38:24–6. doi: 10.1038/ng1718. [DOI] [PubMed] [Google Scholar]
- [149].Geller LN, Potter H. Chromosome missegregation and trisomy 21 mosaicism in Alzheimer’s disease. Neurobiol Dis. 1999;6:167–79. doi: 10.1006/nbdi.1999.0236. [DOI] [PubMed] [Google Scholar]
- [150].Migliore L, Testa A, Scarpato R, Pavese N, Petrozzi L, Bonuccelli U. Spontaneous and induced aneuploidy in peripheral blood lymphocytes of patients with Alzheimer’s disease. Hum Genet. 1997;101:299–305. doi: 10.1007/s004390050632. [DOI] [PubMed] [Google Scholar]
- [151].Migliore L, Botto N, Scarpato R, Petrozzi L, Cipriani G, Bonuccelli U. Preferential occurrence of chromosome 21 malsegregation in peripheral blood lymphocytes of Alzheimer disease patients. Cytogenet Cell Genet. 1999;87:41–6. doi: 10.1159/000015389. [DOI] [PubMed] [Google Scholar]
- [152].Migliore L, Coppede F, Fenech M, Thomas P. Association of micronucleus frequency with neurodegenerative diseases. Mutagenesis. 2011;26:85–92. doi: 10.1093/mutage/geq067. [DOI] [PubMed] [Google Scholar]
- [153].Westra JW, Barral S, Chun J. A reevaluation of tetraploidy in the Alzheimer’s disease brain. Neurodegener Dis. 2009;6:221–9. doi: 10.1159/000236901. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [154].Mosch B, Morawski M, Mittag A, Lenz D, Tarnok A, Arendt T. Aneuploidy and DNA replication in the normal human brain and Alzheimer’s disease. J Neurosci. 2007;27:6859–67. doi: 10.1523/JNEUROSCI.0379-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [155].Copani A, Caraci F, Hoozemans JJ, Calafiore M, Sortino MA, Nicoletti F. The nature of the cell cycle in neurons: focus on a “non-canonical” pathway of DNA replication causally related to death. Biochim Biophys Acta. 2007;1772:409–12. doi: 10.1016/j.bbadis.2006.10.016. [DOI] [PubMed] [Google Scholar]
- [156].Yang Y, Geldmacher DS, Herrup K. DNA replication precedes neuronal cell death in Alzheimer’s disease. J Neurosci. 2001;21:2661–8. doi: 10.1523/JNEUROSCI.21-08-02661.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [157].Yurov YB, Vorsanova SG, Iourov IY. The DNA replication stress hypothesis of Alzheimer’s disease. ScientificWorldJournal. 2011;11:2602–12. doi: 10.1100/2011/625690. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [158].McShea A, Harris PL, Webster KR, Wahl AF, Smith MA. Abnormal expression of the cell cycle regulators P16 and CDK4 in Alzheimer’s disease. Am J Pathol. 1997;150:1933–9. [PMC free article] [PubMed] [Google Scholar]
- [159].Vincent I, Jicha G, Rosado M, Dickson DW. Aberrant expression of mitotic cdc2/cyclin B1 kinase in degenerating neurons of Alzheimer’s disease brain. J Neurosci. 1997;17:3588–98. doi: 10.1523/JNEUROSCI.17-10-03588.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [160].Vincent I, Rosado M, Davies P. Mitotic mechanisms in Alzheimer’s disease? J Cell Biol. 1996;132:413–25. doi: 10.1083/jcb.132.3.413. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [161].Nagy Z, Esiri MM, Cato AM, Smith AD. Cell cycle markers in the hippocampus in Alzheimer’s disease. Acta Neuropathol (Berl) 1997;94:6–15. doi: 10.1007/s004010050665. [DOI] [PubMed] [Google Scholar]
- [162].Copani A, Hoozemans JJ, Caraci F, Calafiore M, Van Haastert ES, Veerhuis R, et al. DNA polymerase-beta is expressed early in neurons of Alzheimer’s disease brain and is loaded into DNA replication forks in neurons challenged with beta-amyloid. J Neurosci. 2006;26:10949–57. doi: 10.1523/JNEUROSCI.2793-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [163].Hernandez-Ortega K, Ferrera P, Arias C. Sequential expression of cell-cycle regulators and Alzheimer’s disease-related proteins in entorhinal cortex after hippocampal excitotoxic damage. J Neurosci Res. 2007;85:1744–51. doi: 10.1002/jnr.21301. [DOI] [PubMed] [Google Scholar]
- [164].Yang Y, Mufson EJ, Herrup K. Neuronal cell death is preceded by cell cycle events at all stages of Alzheimer’s disease. J Neurosci. 2003;23:2557–63. doi: 10.1523/JNEUROSCI.23-07-02557.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [165].Lopes JP, Oliveira CR, Agostinho P. Cdk5 acts as a mediator of neuronal cell cycle re-entry triggered by amyloid-beta and prion peptides. Cell Cycle. 2009;8:97–104. doi: 10.4161/cc.8.1.7506. [DOI] [PubMed] [Google Scholar]
- [166].Esteras N, Bartolome F, Alquezar C, Antequera D, Munoz U, Carro E, et al. Altered cell cycle-related gene expression in brain and lymphocytes from a transgenic mouse model of Alzheimer’s disease [amyloid precursor protein/presenilin 1 (PS1)] Eur J Neurosci. 2012;36:2609–18. doi: 10.1111/j.1460-9568.2012.08178.x. [DOI] [PubMed] [Google Scholar]
- [167].Sherr CJ. G1 phase progression: cycling on cue. Cell. 1994;79:551–5. doi: 10.1016/0092-8674(94)90540-1. [DOI] [PubMed] [Google Scholar]
- [168].Zhu X, Lee HG, Perry G, Smith MA. Alzheimer disease, the two-hit hypothesis: an update. Biochim Biophys Acta. 2007;1772:494–502. doi: 10.1016/j.bbadis.2006.10.014. [DOI] [PubMed] [Google Scholar]
- [169].Ogawa O, Lee HG, Zhu X, Raina A, Harris PL, Castellani RJ, et al. Increased p27, an essential component of cell cycle control, in Alzheimer’s disease. Aging Cell. 2003;2:105–10. doi: 10.1046/j.1474-9728.2003.00042.x. [DOI] [PubMed] [Google Scholar]
- [170].Bonda DJ, Bajic VP, Spremo-Potparevic B, Casadesus G, Zhu X, Smith MA, et al. Review: cell cycle aberrations and neurodegeneration. Neuropathol Appl Neurobiol. 2010;36:157–63. doi: 10.1111/j.1365-2990.2010.01064.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [171].Zhu X, Raina AK, Perry G, Smith MA. Alzheimer’s disease: the two-hit hypothesis. Lancet Neurol. 2004;3:219–26. doi: 10.1016/S1474-4422(04)00707-0. [DOI] [PubMed] [Google Scholar]
- [172].Schubert D, Cole G, Saitoh T, Oltersdorf T. Amyloid beta protein precursor is a mitogen. Biochem Biophys Res Commun. 1989;162:83–8. doi: 10.1016/0006-291x(89)91965-7. [DOI] [PubMed] [Google Scholar]
- [173].Schindowski K, Belarbi K, Bretteville A, Ando K, Buee L. Neurogenesis and cell cycle-reactivated neuronal death during pathogenic tau aggregation. Genes Brain Behav. 2008;7(Suppl 1):92–100. doi: 10.1111/j.1601-183X.2007.00377.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [174].Granic A, Padmanabhan J, Norden M, Potter H. Alzheimer Abeta peptide induces chromosome mis-segregation and aneuploidy, including trisomy 21: requirement for tau and APP. Mol Biol Cell. 2010;21:511–20. doi: 10.1091/mbc.E09-10-0850. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [175].Borysov SI, Granic A, Padmanabhan J, Walczak CE, Potter H. Alzheimer Abeta disrupts the mitotic spindle and directly inhibits mitotic microtubule motors. Cell Cycle. 2011;10:1397–410. doi: 10.4161/cc.10.9.15478. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [176].Judge M, Hornbeck L, Potter H, Padmanabhan J. Mitosis-specific phosphorylation of amyloid precursor protein at threonine 668 leads to its altered processing and association with centrosomes. Mol Neurodegener. 2011;6:80. doi: 10.1186/1750-1326-6-80. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [177].Li J, Xu M, Zhou H, Ma J, Potter H. Alzheimer presenilins in the nuclear membrane, interphase kinetochores, and centrosomes suggest a role in chromosome segregation. Cell. 1997;90:917–27. doi: 10.1016/s0092-8674(00)80356-6. [DOI] [PubMed] [Google Scholar]
- [178].Boeras DI, Granic A, Padmanabhan J, Crespo NC, Rojiani AM, Potter H. Alzheimer’s presenilin 1 causes chromosome missegregation and aneuploidy. Neurobiol Aging. 2008;29:319–28. doi: 10.1016/j.neurobiolaging.2006.10.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [179].Zekanowski C, Wojda U. Aneuploidy, chromosomal missegregation, and cell cycle reentry in Alzheimer’s disease. Acta Neurobiol Exp (Wars) 2009;69:232–53. doi: 10.55782/ane-2009-1748. [DOI] [PubMed] [Google Scholar]
- [180].Nunomura A, Perry G, Aliev G, Hirai K, Takeda A, Balraj EK, et al. Oxidative damage is the earliest event in Alzheimer disease. J Neuropathol Exp Neurol. 2001;60:759–67. doi: 10.1093/jnen/60.8.759. [DOI] [PubMed] [Google Scholar]
- [181].Nunomura A, Perry G, Pappolla MA, Friedland RP, Hirai K, Chiba S, et al. Neuronal oxidative stress precedes amyloid-beta deposition in Down syndrome. J Neuropathol Exp Neurol. 2000;59:1011–7. doi: 10.1093/jnen/59.11.1011. [DOI] [PubMed] [Google Scholar]
- [182].Smith MA, Perry G, Richey PL, Sayre LM, Anderson VE, Beal MF, et al. Oxidative damage in Alzheimer’s. Nature. 1996;382:120–1. doi: 10.1038/382120b0. [DOI] [PubMed] [Google Scholar]
- [183].Smith MA, Hirai K, Hsiao K, Pappolla MA, Harris PL, Siedlak SL, et al. Amyloid-beta deposition in Alzheimer transgenic mice is associated with oxidative stress. J Neurochem. 1998;70:2212–5. doi: 10.1046/j.1471-4159.1998.70052212.x. [DOI] [PubMed] [Google Scholar]
- [184].Perry G, Smith MA. Is oxidative damage central to the pathogenesis of Alzheimer disease? Acta Neurol Belg. 1998;98:175–9. [PubMed] [Google Scholar]
- [185].Taupin P. Aging and neurogenesis, a lesion from Alzheimer’s disease. Aging Dis. 2010;1:158–68. [PMC free article] [PubMed] [Google Scholar]
- [186].Farmer AE, McGuffin P, Gottesman II. Twin concordance for DSM-III schizophrenia. Scrutinizing the validity of the definition. Arch Gen Psychiatry. 1987;44:634–41. doi: 10.1001/archpsyc.1987.01800190054009. [DOI] [PubMed] [Google Scholar]
- [187].Rao DC, Morton NE, Gottesman II, Lew R. Path analysis of qualitative data on pairs of relatives: application to schizophrenia. Hum Hered. 1981;31:325–33. doi: 10.1159/000153233. [DOI] [PubMed] [Google Scholar]
- [188].Mulle JG. Schizophrenia genetics: progress, at last. Curr Opin Genet Dev. 2012;22:238–44. doi: 10.1016/j.gde.2012.02.011. [DOI] [PubMed] [Google Scholar]
- [189].O’Donovan MC, Craddock N, Norton N, Williams H, Peirce T, Moskvina V, et al. Identification of loci associated with schizophrenia by genome-wide association and follow-up. Nat Genet. 2008;40:1053–5. doi: 10.1038/ng.201. [DOI] [PubMed] [Google Scholar]
- [190].Stefansson H, Ophoff RA, Steinberg S, Andreassen OA, Cichon S, Rujescu D, et al. Common variants conferring risk of schizophrenia. Nature. 2009;460:744–7. doi: 10.1038/nature08186. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [191].Stefansson H, Rujescu D, Cichon S, Pietilainen OP, Ingason A, Steinberg S, et al. Large recurrent microdeletions associated with schizophrenia. Nature. 2008;455:232–6. doi: 10.1038/nature07229. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [192].McClellan JM, Susser E, King MC. Schizophrenia: a common disease caused by multiple rare alleles. Br J Psychiatry. 2007;190:194–9. doi: 10.1192/bjp.bp.106.025585. [DOI] [PubMed] [Google Scholar]
- [193].Tam GW, Redon R, Carter NP, Grant SG. The role of DNA copy number variation in schizophrenia. Biol Psychiatry. 2009;66:1005–12. doi: 10.1016/j.biopsych.2009.07.027. [DOI] [PubMed] [Google Scholar]
- [194].Sahoo T, Theisen A, Rosenfeld JA, Lamb AN, Ravnan JB, Schultz RA, et al. Copy number variants of schizophrenia susceptibility loci are associated with a spectrum of speech and developmental delays and behavior problems. Genet Med. 2011;13:868–80. doi: 10.1097/GIM.0b013e3182217a06. [DOI] [PubMed] [Google Scholar]
- [195].McCarthy SE, Makarov V, Kirov G, Addington AM, McClellan J, Yoon S, et al. Microduplications of 16p11.2 are associated with schizophrenia. Nat Genet. 2009;41:1223–7. doi: 10.1038/ng.474. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [196].Rare chromosomal deletions and duplications increase risk of schizophrenia. Nature. 2008;455:237–41. doi: 10.1038/nature07239. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [197].Williams NM, O’Donovan MC, Owen MJ. Chromosome 22 deletion syndrome and schizophrenia. Int Rev Neurobiol. 2006;73:1–27. doi: 10.1016/S0074-7742(06)73001-X. [DOI] [PubMed] [Google Scholar]
- [198].Murphy KC, Jones LA, Owen MJ. High rates of schizophrenia in adults with velo-cardio-facial syndrome. Arch Gen Psychiatry. 1999;56:940–5. doi: 10.1001/archpsyc.56.10.940. [DOI] [PubMed] [Google Scholar]
- [199].Vassos E, Collier DA, Holden S, Patch C, Rujescu D, St Clair D, et al. Penetrance for copy number variants associated with schizophrenia. Hum Mol Genet. 2010;19:3477–81. doi: 10.1093/hmg/ddq259. [DOI] [PubMed] [Google Scholar]
- [200].DeLisi LE, Friedrich U, Wahlstrom J, Boccio-Smith A, Forsman A, Eklund K, et al. Schizophrenia and sex chromosome anomalies. Schizophr Bull. 1994;20:495–505. doi: 10.1093/schbul/20.3.495. [DOI] [PubMed] [Google Scholar]
- [201].Bassett AS, Chow EW, Weksberg R. Chromosomal abnormalities and schizophrenia. Am J Med Genet. 2000;97:45–51. doi: 10.1002/(sici)1096-8628(200021)97:1<45::aid-ajmg6>3.0.co;2-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [202].Yurov YB, Vostrikov VM, Vorsanova SG, Monakhov VV, Iourov IY. Multicolor fluorescent in situ hybridization on post-mortem brain in schizophrenia as an approach for identification of low-level chromosomal aneuploidy in neuropsychiatric diseases. Brain Dev. 2001;23(Suppl 1):S186–90. doi: 10.1016/s0387-7604(01)00363-1. [DOI] [PubMed] [Google Scholar]
- [203].Yurov YB, Iourov IY, Vorsanova SG, Demidova IA, Kravetz VS, Beresheva AK, et al. The schizophrenia brain exhibits low-level aneuploidy involving chromosome 1. Schizophr Res. 2008;98:139–47. doi: 10.1016/j.schres.2007.07.035. [DOI] [PubMed] [Google Scholar]
- [204].Fombonne E. The prevalence of autism. JAMA. 2003;289:87–9. doi: 10.1001/jama.289.1.87. [DOI] [PubMed] [Google Scholar]
- [205].Chakrabarti S, Fombonne E. Pervasive developmental disorders in preschool children: confirmation of high prevalence. Am J Psychiatry. 2005;162:1133–41. doi: 10.1176/appi.ajp.162.6.1133. [DOI] [PubMed] [Google Scholar]
- [206].Vorstman JA, Staal WG, van Daalen E, van Engeland H, Hochstenbach PF, Franke L. Identification of novel autism candidate regions through analysis of reported cytogenetic abnormalities associated with autism. Mol Psychiatry. 2006;11:1, 18–28. doi: 10.1038/sj.mp.4001781. [DOI] [PubMed] [Google Scholar]
- [207].Martin CL, Ledbetter DH. Autism and cytogenetic abnormalities: solving autism one chromosome at a time. Curr Psychiatry Rep. 2007;9:141–7. doi: 10.1007/s11920-007-0084-9. [DOI] [PubMed] [Google Scholar]
- [208].Sebat J, Lakshmi B, Malhotra D, Troge J, Lese-Martin C, Walsh T, et al. Strong association of de novo copy number mutations with autism. Science. 2007;316:445–9. doi: 10.1126/science.1138659. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [209].Marshall CR, Noor A, Vincent JB, Lionel AC, Feuk L, Skaug J, et al. Structural variation of chromosomes in autism spectrum disorder. Am J Hum Genet. 2008;82:477–88. doi: 10.1016/j.ajhg.2007.12.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [210].Allen-Brady K, Robison R, Cannon D, Varvil T, Villalobos M, Pingree C, et al. Genome-wide linkage in Utah autism pedigrees. Mol Psychiatry. 2010;15:1006–15. doi: 10.1038/mp.2009.42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [211].Liu J, Nyholt DR, Magnussen P, Parano E, Pavone P, Geschwind D, et al. A genomewide screen for autism susceptibility loci. Am J Hum Genet. 2001;69:327–40. doi: 10.1086/321980. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [212].Veenstra-VanderWeele J, Cook EH., Jr. Molecular genetics of autism spectrum disorder. Mol Psychiatry. 2004;9:819–32. doi: 10.1038/sj.mp.4001505. [DOI] [PubMed] [Google Scholar]
- [213].Dykens EM, Sutcliffe JS, Levitt P. Autism and 15q11-q13 disorders: behavioral, genetic, and pathophysiological issues. Ment Retard Dev Disabil Res Rev. 2004;10:284–91. doi: 10.1002/mrdd.20042. [DOI] [PubMed] [Google Scholar]
- [214].Szatmari P, Paterson AD, Zwaigenbaum L, Roberts W, Brian J, Liu XQ, et al. Mapping autism risk loci using genetic linkage and chromosomal rearrangements. Nat Genet. 2007;39:319–28. doi: 10.1038/ng1985. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [215].Mefford HC, Sharp AJ, Baker C, Itsara A, Jiang Z, Buysse K, et al. Recurrent rearrangements of chromosome 1q21.1 and variable pediatric phenotypes. New Engl J Med. 2008;359:1685–99. doi: 10.1056/NEJMoa0805384. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [216].Liang JS, Shimojima K, Ohno K, Sugiura C, Une Y, Yamamoto T. A newly recognised microdeletion syndrome of 2p15-16.1 manifesting moderate developmental delay, autistic behaviour, short stature, microcephaly, and dysmorphic features: a new patient with 3.2 Mb deletion. J Med Genet. 2009;46:645–7. doi: 10.1136/jmg.2008.059220. [DOI] [PubMed] [Google Scholar]
- [217].Muhle R, Trentacoste SV. Rapin I. The genetics of autism. Pediatrics. 2004;113:e472–86. doi: 10.1542/peds.113.5.e472. [DOI] [PubMed] [Google Scholar]
- [218].Shinawi M, Sahoo T, Maranda B, Skinner SA, Skinner C, Chinault C, et al. 11p14.1 microdeletions associated with ADHD, autism, developmental delay, and obesity. Am J Med Genet A. 2011;155A:1272–80. doi: 10.1002/ajmg.a.33878. [DOI] [PubMed] [Google Scholar]
- [219].Weiss LA, Shen Y, Korn JM, Arking DE, Miller DT, Fossdal R, et al. Association between microdeletion and microduplication at 16p11.2 and autism. New Engl J Med. 2008;358:667–75. doi: 10.1056/NEJMoa075974. [DOI] [PubMed] [Google Scholar]
- [220].Hanson E, Nasir RH, Fong A, Lian A, Hundley R, Shen Y, et al. Cognitive and behavioral characterization of 16p11.2 deletion syndrome. J Dev Behav Pediatr. 2010;31:649–57. doi: 10.1097/DBP.0b013e3181ea50ed. [DOI] [PubMed] [Google Scholar]
- [221].Moreno-De-Luca D, Mulle JG, Kaminsky EB, Sanders SJ, Myers SM, Adam MP, et al. Deletion 17q12 is a recurrent copy number variant that confers high risk of autism and schizophrenia. Am J Hum Genet. 2010;87:618–30. doi: 10.1016/j.ajhg.2010.10.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [222].Antshel KM, Aneja A, Strunge L, Peebles J, Fremont WP, Stallone K, et al. Autistic spectrum disorders in velo-cardio facial syndrome (22q11.2 deletion) J Autism Dev Disord. 2007;37:1776–86. doi: 10.1007/s10803-006-0308-6. [DOI] [PubMed] [Google Scholar]
- [223].Reddy KS. Cytogenetic abnormalities and fragile-X syndrome in Autism Spectrum Disorder. BMC Med Genet. 2005;6:3. doi: 10.1186/1471-2350-6-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [224].van Rijn S, Swaab H, Aleman A, Kahn RS. Social behavior and autism traits in a sex chromosomal disorder: Klinefelter (47XXY) syndrome. J Autism Dev Disord. 2008;38:1634–41. doi: 10.1007/s10803-008-0542-1. [DOI] [PubMed] [Google Scholar]
- [225].Yurov YB, Vorsanova SG, Iourov IY, Demidova IA, Beresheva AK, Kravetz VS, et al. Unexplained autism is frequently associated with low-level mosaic aneuploidy. J Med Genet. 2007;44:521–5. doi: 10.1136/jmg.2007.049312. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [226].Zhao X, Leotta A, Kustanovich V, Lajonchere C, Geschwind DH, Law K, et al. A unified genetic theory for sporadic and inherited autism. Proc Natl Acad Sci U S A. 2007;104:12831–6. doi: 10.1073/pnas.0705803104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [227].Geschwind DH, Sowinski J, Lord C, Iversen P, Shestack J, Jones P, et al. The autism genetic resource exchange: a resource for the study of autism and related neuropsychiatric conditions. Am J Hum Genet. 2001;69:463–6. doi: 10.1086/321292. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [228].Allen-Brady K, Cannon D, Robison R, McMahon WM, Coon H. A unified theory of autism revisited: linkage evidence points to chromosome X using a high-risk subset of AGRE families. Autism Res. 2010;3:47–52. doi: 10.1002/aur.119. [DOI] [PubMed] [Google Scholar]
- [229].Pampanos A, Volaki K, Kanavakis E, Papandreou O, Youroukos S, Thomaidis L, et al. A substitution involving the NLGN4 gene associated with autistic behavior in the Greek population. Genet Test Mol Biomarkers. 2009;13:611–5. doi: 10.1089/gtmb.2009.0005. [DOI] [PubMed] [Google Scholar]
- [230].Hanks S, Coleman K, Reid S, Plaja A, Firth H, Fitzpatrick D, et al. Constitutional aneuploidy and cancer predisposition caused by biallelic mutations in BUB1B. Nat Genet. 2004;36:1159–61. doi: 10.1038/ng1449. [DOI] [PubMed] [Google Scholar]
- [231].Callier P, Faivre L, Cusin V, Marle N, Thauvin-Robinet C, Sandre D, et al. Microcephaly is not mandatory for the diagnosis of mosaic variegated aneuploidy syndrome. Am J Med Genet A. 2005;137:204–7. doi: 10.1002/ajmg.a.30783. [DOI] [PubMed] [Google Scholar]
- [232].Micale MA, Schran D, Emch S, Kurczynski TW, Rahman N, Van Dyke DL. Mosaic variegated aneuploidy without microcephaly: implications for cytogenetic diagnosis. Am J Med Genet A. 2007;143A:1890–3. doi: 10.1002/ajmg.a.31848. [DOI] [PubMed] [Google Scholar]
- [233].Matsuura S, Matsumoto Y, Morishima K, Izumi H, Matsumoto H, Ito E, et al. Monoallelic BUB1B mutations and defective mitotic-spindle checkpoint in seven families with premature chromatid separation (PCS) syndrome. Am J Med Genet A. 2006;140:358–67. doi: 10.1002/ajmg.a.31069. [DOI] [PubMed] [Google Scholar]
- [234].Kawame H, Sugio Y, Fuyama Y, Hayashi Y, Suzuki H, Kurosawa K, et al. Syndrome of microcephaly, Dandy-Walker malformation, and Wilms tumor caused by mosaic variegated aneuploidy with premature centromere division (PCD): report of a new case and review of the literature. J Hum Genet. 1999;44:219–24. doi: 10.1007/s100380050147. [DOI] [PubMed] [Google Scholar]
- [235].Jacquemont S, Boceno M, Rival JM, Mechinaud F, David A. High risk of malignancy in mosaic variegated aneuploidy syndrome. Am J Med Genet. 2002;109:17–21. doi: 10.1002/ajmg.10281. discussion 16. [DOI] [PubMed] [Google Scholar]
- [236].Limwongse C, Schwartz S, Bocian M, Robin NH. Child with mosaic variegated aneuploidy and embryonal rhabdomyosarcoma. Am J Med Genet. 1999;82:20–4. doi: 10.1002/(sici)1096-8628(19990101)82:1<20::aid-ajmg4>3.0.co;2-5. [DOI] [PubMed] [Google Scholar]