Skip to main content
Current Genomics logoLink to Current Genomics
. 2010 Sep;11(6):387–396. doi: 10.2174/138920210793176065

Somatic Genome Variations in Health and Disease

IY Iourov 1,2,3,*, SG Vorsanova 1,2,3, YB Yurov 1,2
PMCID: PMC3018718  PMID: 21358982

Abstract

It is hard to imagine that all the cells of the human organism (about 1014) share identical genome. Moreover, the number of mitoses (about 1016) required for the organism’s development and maturation during ontogeny suggests that at least a proportion of them could be abnormal leading, thereby, to large-scale genomic alterations in somatic cells. Experimental data do demonstrate such genomic variations to exist and to be involved in human development and interindividual genetic variability in health and disease. However, since current genomic technologies are mainly based on methods, which analyze genomes from a large pool of cells, intercellular or somatic genome variations are significantly less appreciated in modern bioscience. Here, a review of somatic genome variations occurring at all levels of genome organization (i.e. DNA sequence, subchromosomal and chromosomal) in health and disease is presented. Looking through the available literature, it was possible to show that the somatic cell genome is extremely variable. Additionally, being mainly associated with chromosome or genome instability (most commonly manifesting as aneuploidy), somatic genome variations are involved in pathogenesis of numerous human diseases. The latter mainly concerns diseases of the brain (i.e. autism, schizophrenia, Alzheimer’s disease) and immune system (autoimmune diseases), chromosomal and some monogenic syndromes, cancers, infertility and prenatal mortality. Taking into account data on somatic genome variations and chromosome instability, it becomes possible to show that related processes can underlie non-malignant pathology such as (neuro)degeneration or other local tissue dysfunctions. Together, we suggest that detection and characterization of somatic genome behavior and variations can provide new opportunities for human genome research and genetics.

Keywords: Somatic genome variations, aneuploidy, chromosome instability, genomic instability, disease.

INTRODUCTION

The human organism consists of about 1014 cells of 210 different types that originate from one zygote and are the result of about 1016 mitoses (approximately 45 cellular generations). Moreover, the large number of cellular divisions is required to maintain relatively stable amount of cells in a human body to cover each day’s loss of more than several tens of millions of cells. These numbers make clear that such an “amount of processes” cannot be identically reproduced and, therefore, all the cells of an organism are unlikely to possess identical genomes.

It appears that the most critical period for somatic mutations to occur is early embryonic development [1, 2]. Characterized by the logarithmic increase in cells (the most dramatic increase of cell numbers in human ontogeny) [2], human embryos are thought to exhibit increased levels of mitotic mutations [3-5]. This is experimentally confirmed by molecular cytogenetic studies of embryonic and fetal cells, which demonstrate high rates of aneuploidy due to mitotic errors correlated with high cell division rate [5, 6]. Therefore, genetically altered cells produced during this ontogenetic period form a basis for organism dysfunction at the following developmental stages [3, 4, 7]. Nonetheless, there is still a possibility that somatic genome variations (SGV) lack adverse effect due to natural selection and clearance of abnormal cells [5, 4, 7, 8].

The genome of a cell is supposed to experience 104-105 of DNA lesions per day. This is another source (exogenous source) for cellular genome to change and, if remains unrepaired (uncleared), such genomic variations give rise to pathogenic processes (i.e. cancerization) [9]. Although this is a likely process for diseases caused by SGV produced through either genomic instability (GIN) or chromosome instability (CIN) [7-12], it is supposed to be an underlying mechanism of human aging [8, 11].

Despite of numerous attempts to highlight the role of SGV [2-8, 10, 12-22], related phenomena remain largely underappreciated in current biomedical literature. This suggests that an additional attention to SGV is required. Hence, a review of SGV might help to define the contribution to human interindividual diversity in health and disease.

NATURAL SGV

Since benign genomic variations of somatic genome remain to be poorly described, it is hard to assess the effect of SGV on the non-pathogenic diversity. Thus, no less than 12% of the human genome encompassing disease-associated loci is diversified between two individuals [23]. Although it is difficult to extrapolate these data to cell populations, it can be considered as an indirect evidence for cellular genome to change in a related manner. Fortunately, there are molecular cytogenetic data on SGV manifested at chromosomal level (structural rearrangements, aneuploidy and polyploidy) in early prenatal development. As to other types of genomic variations, including single-base DNA changes, DNA sequence deletion/duplications/inversions, repeat expansions, transposition of mobile DNA elements, copy number variations (CNV), chromosomal miscrodeletions/microduplications (for more details see reviews [15, 24]), their incidence among human fetuses remain largely unknown. SGV detected after birth (non-affected individuals) are mainly referred to low-level mosaic aneuploidy [3, 4, 7]. Mosaic structural genomic rearrangements at chromosomal level are also reported, being, however occasionally detected. Additionally, the best documented SGV are tissue-specific variations of chromosome numbers (aneuploidy) and CNV. Table 1 gathers the data on SGV in normal human tissues (cell types).

Table 1.

SGV in Normal Human Tissues

Tissue/Cell type Type of SGV Description Key Refs
Preimplantation embryos Aneuploidy 15-91% of samples (mean is about 50%) [6, 25, 26]
Structural rearrangements, aneuploidy, CNV*, segmental duplications, uniparental disomy >90 %of samples (83 % — aneuploidy) [6]
Embryos/Fetuses (7-12 weeks)
Cytotrophoblasts Aneuploidy 20-60% of cells [27]
Brain Aneuploidy 1.45% of cells ^ [5, 28]
Chorionic villi Aneuploidy 0.98% of cells ^ [5]
Skin Aneuploidy 0.82% of cells^ [5]
Ovarian tissue Trisomy of chromosome 21 (aneuploidy) Statistically significant increase of aneuploid cells [29]
Prenatal diagnosis: CVS** or Amniocentesis
Amniocytes Aneuploidy 0.25% of samples [30]
Chorionic villi/Placenta Aneuploidy 1-2% of samples [31, 32]
Newborns/Children
Blood lymphocytes Aneuploidy >0.1% (clinical population?) [4, 33, 34]
Blood lymphocytes Aneuploidy 0.73% (autosomes) and 1.11% (chromosome X) of cells — unaffected population^ [35]
Blood lymphocytes Structural rearrangements 0.01% (clinical population?) [36]
Adults (middle age)
Blood lymphocytes Aneuploidy 1-3% of cells^ [37-39]
Blood lymphocytes Structural rearrangements 0.6% of cells [40]
Skin fibroblasts Aneuploidy 2.2% of cells^ [41]
Liver Aneuploidy 3% of cells^ [42]
Brain Aneuploidy 0.3-0.9% of cells^ [10, 28, 43-45]
Brain CNV* Tissue-specific CNV; amount of cells and percentage of samples was not available [46]
Skin
Heart
Kidney
Liver
T-lymphocytes Subtle structural rearrangements or CNV Tissue-specific mosaicism probably originating from developmental chromosome instability [47]
Imortalized B lymphoblastoid cells
Skin fibroblasts
Adults (aged individuals)
Blood lymphocytes Aneuploidy 1-2% (autosomes) and 4-7% (chromosome X) of cells^ [37, 39]
Skin fibroblasts Aneuploidy 4.4% of cells^ [41]
Brain Aneuploidy 0.3-0.9% of cells^ [10, 28, 43-45]
*

— copy number variations;

**

— chorionic villus sampling;

^

— per chromosome.

The essential problem surrounding the evaluation of SGV is a technological one (for more details see [3, 4, 7, 15, 18, 22, 24]). In other words, some tissues or developmental stages were evaluated using single-cell high-resolution molecular cytogenetic techniques, whereas others were not [22]. Therefore, it is hard to compare different data on SGV. Nevertheless, preimplantation embryos exhibit high rates of SGV manifested at chromosomal (microscopic and submicroscopic) level including aneuploidy, gross structural genomic rearrangements, CNV, segmental duplications. In total, it is estimated that almost 90% of samples have cells with different genomes [6, 25, 26]. The intercellular rate of variations (percentage of abnormal cells) is uninformative because of small amount of cells at this developmental stage [3, 4]. At the next stages of prenatal development, a lesser frequency of SGV is observed, being, still, appreciable and affecting up to 30% of fetuses (aneuploidy) [4, 5, 8]. This is observed in extraembrionic tissues [5, 27] and is suggested to play a key role in normal human placentation [27]. Additionally, no less than 30-35% of cells of the developing human brain and 20% of fetal skin are aneuploid [5, 28]. Finally, fetal ovarian tissues demonstrate a significant increase of mosaic trisomy of chromosome 21 [29]. Further periods of human intrauterine development ascertained through prenatal diagnosis (chorionic villus sampling (CVS) and amniocentesis) show a small rate of SGV [30]. However, it is to note that these periods are rarely evaluated by molecular cytogenetic techniques (singular case-reports only), which are essential for accurate SGV detection [3, 22]. An additional issue of such studies is description of another example of apparently benign tissue-specific SGV in human fetuses referred to as placental mosaicism [31, 32]. Together, SGV appear to be mainly formed during prenatal development and have the potential to give rise to intercellular diversity after birth in health and disease.

Newborns were not thoroughly evaluated in terms of SGV. Furthermore, the only available data on large genomic variations (chromosomal abnormalities and heteromorphism) can be only acquired from papers describing banding studies performed in the end of 70s or beginning of 80s [4, 33, 34]. Chromosomal mosaicism detected by banding cytogenetics in newborns is less frequent than 0.1% [33, 34]. Molecular cytogenetic evaluations of control group in a survey of SGV in autism showed rates of mosaic aneuploidy in blood lymphocytes as 0.73% (autosomes) and 1.11% (chromosome X) [35]. Mosaic structural chromosome rearrangements are extremely rare and are supposed to be detected in a clinical population only [36]. Middle age adults exhibit 1-3% of aneuploid cells in tissues composed of mitotically active cells and less than 1% in the adult human brain, which is mainly composed of post-mitotic cells [10, 28, 37-45]. Natural SGV manifesting as structural rearrangements detectable by banding cytogenetics in blood lymphocytes achieve the rate of 0.6% [40]. Mosaic subtle structural genomic rearrangements and CNV can be tissue-specific in presumably unaffected individuals [46, 47]. Aged human tissues are known to be featured by increased rates of SGV essentially manifesting as low-level mosaic aneuploidy [8, 10, 28, 37-39, 41, 43-45]. In conclusion, three main features of natural human SGV may be highlighted: (i) SGV do contribute to human natural (intercellular) genomic variation; (ii) further studies are strongly required to identify incidence and possible effect of SGV on unaffected human tissues; (iii) SGV have different rates at different ontogenetic stages. The latter suggests a role for SGV in human development and aging.

SGV AND DEVELOPMENTAL/AGING PROCESSES

Although involvements of SGV in developmental and aging processes are presented in another review published in this Hot Topic Issue (YB Yurov et al. Ontogenetic variation of the human genome), we found needful to mention briefly related phenomena. This appears to be important for further delineation of the role of SGV in human diseases and mechanisms of SGV formation. Two kinds of fates of abnormal cells formed during early prenatal development are hypothesized: persistence (increase or stability of rates) and clearance (decrease of rates). The former is supposed to represent a mechanism for SGV-associated diseases (i.e. brain diseases, cancers, mosaic chromosome abnormalities), where as the latter is likely to be a normal process aimed at regulation of cellular population size and to protect against aneuploidization or other unfavorable SGV [4, 5, 7, 8, 47, 48]. Similar processes appear to underlie human aging, including diseases of pathological/accelerated aging [8, 10, 18, 45]. In sum, this suggests that SGV formed during prenatal development are probably responsible for human prenatal mortality and postnatal morbidity. However, SGV originating from somatic mutations after birth are likely to be diseases-causing, as well (as exemplified by studying GIN and CIN in cancers).

SGV AND HEREDITARY DISEASES

Genomic variations are determined according to DNA sequence size that is involved in a rearrangement [24]. Numerous studies performed during the last decade were focused on genomic variations at DNA sequence level (gene mutations) [2, 3, 13, 19, 24] and copy number variations (CNV) [2, 3, 6, 15, 16, 23, 24, 46]. In this extent, SGV was continuously studied in monogenic syndromes and diseases associated with CNV [2, 19]. Table 2 summarizes current data on SGV contribution to pathogenesis of hereditary diseases caused by gene mutations and CNV.

Table 2.

SGV and Hereditary Diseases Demonstrating Somatic Gene Mutations or CNV (in Parts Adopted from [2] and [16])

Locus Disease Gene CNV Gene Mutations
1q21.2 Progeria LMNA - +
1q44 Chronic infantile neurologic cutaneous articular CIAS1 - +
2p22p21 Hereditary spastic paraplegia SPG4 - +
2q24 Myoclonic epilepsy SCN1A - +
2q31 Ehlers Danlos Syndrome IV COL3A1 - +
3p25 von-Hippel-Lindau Disease VHL + +
3q13.3q21 Hypocalcemia CASR - +
3q27 EEC p63 - +
4p16.3 Skeletal disorders (syndromes) FGFR3 - +
4p12 Congenital central hypoventilation PHOX2B - +
4q35 Facioscapulohumeral muscular dystrophy D4Z4* + ?
5q13 Infantile spinal muscular atrophy SMN1 - +
6p21 Cleidocranial dysplasia RUNX2 - +
7q22.1 Osteogenesis imperfecta COL1A2 - +
8q12.1 CHARGE syndrome CHD7 ? +
9q22 Loeys-Dietz TGFBR2 - +
11p15.5 Costello syndrome HRAS - +
11p15.1 Neonatal diabetes KCNJ11 - +
12q13 Epidermolysis bullosa simplex KRT5 - +
12q24.1 Phenylketonuria PAH - +
13q14 Retinoblastoma RB + +
14q24.3 Alzheimer disease, early-onset PS1 - +
15q21.1 Marfan FBN1 - +
16p13 Tuberous Sclerosis TSC2 + +
16p13 Rubinstein-Taybi Syndrome CREBBP + ?
17q11 Neurofibromatosis 1 NF1 + +
17q21.31 Osteogenesis imperfecta COL1A1 - +
17q24 Campomelic dysplasia SOX9 + +
22q11.2 Several hereditary syndromes MYH9 + +
Xp22.2p22.1 X-linked hypophosphatemia PHEX - +
Xp22.13 X-linked mental retardation (syndromic/nonsyndromic) ARX - +
Xp21 Duchenne muscular dystrophy DMD + +
Xp21 Chronic granulomatous disease CYBB + +
Xp21.1 Ornithine transcarbamylase deficiency OTC - +
Xp21.1 Retinitis pigmentosa RPGR - +
Xp11.3 Retinitis pigmentosa RP2 - +
Xq11q12 Androgen insensitivity AR - +
Xq26q27.2 Lesch-Nyhan HPRT1 - +
Xq27 Hemophilia B F9 - +
Xq28 Hemophilia A F8 + +
Xq28 Incontinentia pigmenti IKBKG + +
Xq28 Mucopolysaccharidosis II IDS - +
Xq28 Otopalatodigital syndrome FLNA - +
Xq28 Rett syndrome (males and females) and a set of other neurodevelopmental diseases (syndromic/nonsyndromic) MECP2 + +
Xq28 X-linked dyskeratosis congenita DKC1 + +
Xq28 X-linked mental retardation SLC6A8 - +
*

— non-coding DNA sequences (repeats).

It is probable that some somatic CNV encompassing these genes are, as yet, undescribed due to extreme rarity of the these conditions (at least some of these conditions) [2, 19]. Additional important issue of somatic gene mutations and CNV is related to explanation of phenotypic difference between cases of the same syndrome due to different expressivity in cases of SGV [19]. Finally, the list of somatic gene mutations and CNV is far from being complete. Furthermore, some of them appear to be benign in a proportion of cases [2].

SGV AND CHROMOSOME SYNDROMES

The best documented genomic variations are those detected at submicroscopic and microscopic levels (subtle structural genomic rearrangements and chromosomal abnormalities) [3, 15, 24]. As mentioned above, mosaic structural chromosomal rearrangements are rare. There are few population-based cytogenetic studies of these SGV suggesting them to be associated with milder manifestations of the corresponding non-mosaic rearrangement [36]. Mosaic subtle structural chromosome abnormalities (undetectable by banding cytogenetic techniques) are repeatedly reported but the incidence remain to be estimated [47, 49]. Consequently, this review part is primarily focused on numerical chromosome abnormalities (aneuploidy and poliploidy). According to previous review of chromosomal mosaicism [4], mosaic aneuploidy can be divided into three major groups: rare mosaic autosomal aneuploidy (chromosomes 1, 2, 3, 4, 5, 6, 7, 10, 11, 12, 17, and 19); relatively rare mosaic autosomal aneuploidy (chromosomes 14, 15, 16, and 20); frequent mosaic autosomal aneuploidy (chromosomes 8, 9, 13, 18, 21, and 22). Aneuploidy of sex chromosomes is common due to reduced phenotypic effect as to autosomal aneuploidy [3, 4, 15]. The majority of aneuploidy is trisomy or additional sex chromosomes, inasmuch as loss of an autosome leads to intrauterine death at the earliest stages of prenatal development in contrast to loss of chromosomes X and Y [4, 50, 51]. Interestingly, trisomies of chromosomes, rarely involved in aneuploidy in fetuses and liveborn infants (adults), occur at the same rate in preimplantation embryos as mosaic trisomies of other chromosomes [52]. This suggests that mosaic aneuploidy does not possess appreciable effect on the earliest stage of embryonic development. Therefore, the next stages of the development should exhibit high rates of chromosomal mosaicism. This is supported by data on spontaneous abortions, 25% of which are chromosomal mosaics [50]. Additionally, the presence of uniparental disomy in liveborns is considered a confirmation of cleared prenatal mosaicism (confined placental mosaicism) [53]. Chromosomal mosaicism is also associated with asymmetry and skin pigmentary anomalies [54]. For instance, some syndromes featured by congenital asymmetric deformations exhibit unshared distribution of aneuploid or polyploidy cell lines [3, 4, 54].

After birth, mosaic chromosomal abnormalities are essentially identified among individuals with phenotypic manifestation of recognizable aneuploidy (chromosomal) syndromes [3, 4, 15, 55]. However, there are several reports about unaffected individuals with up to 30% of abnormal (aneuploid) cells (reviewed in [3] and [4]). The proportion of mosaics reflects unequal susceptibility of different chromosomes to mitotic non-disjunction [4, 55]. Table 3 shows mosaicism (mitotic non-disjunction) among cases of aneuploidy in humans.

Table 3.

Mosaic Cases Among Common Aneuploidies (in Parts from [3, 4, 15, 55])

Aneuploidy Cases of Mosaicism/Mitotic Non-Disjunction Incidence Disease
Trisomy of chromosome 2 7% unknown
Trisomy of chromosome 7 57%* unknown
Trisomy of chromosome 8 50% * >100 cases reported Trisomy 8
Trisomy of chromosome 13 1% 1:6000-1:29000 Patau syndrome
Trisomy of chromosome 14 8% ~25 cases reported Trisomy 14
Trisomy of chromosome 15 None ~10 cases reported
Trisomy of chromosome 16 None ~10 cases reported
Trisomy of chromosome 18 8% 1:7000 Edwards syndrome
Trisomy of chromosome 21 5% 1:600 Down syndrome
Trisomy of chromosome 22 2%* Cat eye syndrome (?)
Monosomy of chromosome X 38%* 1:2000 (females) Turner sydnrome
Trisomy of chromosome X 20% 1:1000 (females) Trisomy X
47,XXY 9% 1:500 (males) Klinefelter syndrome
47,XYY 16% 1:800 (males) Double Y syndrome
*

— postnatal cases suggested to be all mosaic.

Another well-described examples of SGV associated with chromosomal abnormalities (chromosomal syndromes), are small supernumerary marker chromosomes. Over 50% of cases demonstrating these chromosomal rearrangements can exhibit mosaicism, including tissue-specific forms [4, 56-58]. More precise information about SGV and marker chromosomes is presented in another review of this Hot Topic Issue (T Liehr et al. Somatic mosaicism in cases with small supernumerary marker chromosomes).

SGV demonstrate a diminished clinical effect of chromosome abnormalities [3, 4, 7, 15]. However, cancers, which are all caused by somatic mutations, are primarily associated with cellular (tissular) pathology [11, 13, 14, 17, 20, 41, 42]. Taking into account these facts, a hypothesis suggesting that SGV manifested as somatic chromosomal mutations (the commonest type of SGV) can be a source or a susceptibility factor for complex human diseases was proposed [3, 4, 7, 8, 12, 15, 48].

SGV AND COMPLEX DISEASES

Currently, SGV have been described in individuals with brain diseases (psychiatric and neurodegenerative) [2-4, 7, 15-17, 19, 22, 35, 55, 59-80], autoimmune diseases [81-83], congenital heart diseases [84] and cancer [9, 11, 13-15, 20, 85]. All these data is summarized by Table 4.

Table 4.

SGV in Complex Human Diseases

Diseasee Type of SGV Key Refs
Brain diseases (psychiatric)
Learning disability/Mental retardation Gene mutations, CNV mosaic aneuploidy [ 2-4, 7, 15-17, 19, 22, 55]
Autism Mosaic structural/numerical chromosomal abnormalities:
Partial tetrasomy 3q
Ring chromosome 14
Rearrangements of 15pter-q13.2
Ring chromosome 17
Structural abnormalities + ring chromosome 18
Mosaic deletion 20p
[59-61]
[62]
[63]
[64, 65]
[66]
[67, 68]
[69]
Mosaic aneuploidy (~16% of cases) [35]
Fragile sites [70, 71]
Schizophrenia Mosaic sex chromosome aneuploidy (blood lymphocytes) [72-76]
Low-level mosaic aneuploidy of chromosomes 1, 18 and X in the diseased brain [12, 44]
Fragile sites [7, 77]
Brain diseases (neurodegenerative)
Alzheimer’s disease Gene mutations [78]
Mosaic aneuploidy of chromosome 21 in the diseased brain [10]
Huntington’s disease Gene mutations (trinucleotide repeat expansion) including brain-specific mutations [79]
Friedreich ataxia Gene mutations (trinucleotide repeat expansion) [80]
Ataxia-telangiectasia Mosaic aneuploidy and chromosome 14-specific breaks/additional rearranged chromosomes [45]
Autoimmune diseases
Primary immune deficiencies Revertant somatic mosaicism [81]
Primary biliary cirrhosis Mosaic monosomy of chromosome X [82]
Systemic sclerosis Mosaic monosomy of chromosome X [83]
Autoimmune thyroid disease
Heart disease
Congenital heart diseases Gene mutations [84]
Chromosomal abnormalities/syndromes (?) [3, 17, 55,]
Cancers
Almost all types of cancers Almost all cancers are caused by different types of SGV including aneuploidy/polyploidy; balanced and unbalanced structural chromosomal/genomic (subtle and gross) rearrangements; gene amplifications; telomere shortening; microsatellite instability; gene mutations; [9, 11, 13-15, 20, 85,]

Theoretically, any mutation can be somatic leading, thereby, to a disease [3]. However, since genomic variations are classically thought to be a result of germline mutations or genomic rearrangements [13, 86], SGV contributions to human morbidity is poorly appreciated. In contrast, complex diseases appear to be likely associated with somatic mutations as to monogenic syndromes and genomic disorders, being commonly associated with tissue-specific (or even “subtissue-specific”) pathology without any additional dysfunctions in other unaffected tissues [3, 4, 7, 12]. Moreover, SGV are able to explain some important features of complex disease-causing genetic alterations such as environmental effects or specific male-to-female ratios [4, 8, 82, 87-89]. Finally, to get an integral view of how SGV and other disease-causing processes interplay with each other, formation mechanisms of somatic mosaicism are to be established.

SGV FORMATION MECHANISMS

Although formation of somatic mutations (aneuploidy and polyploidy) was the major focus of numerous studies in fields of cell biology, genetics and oncology, it is still incompletely understood. A number of neonatal mosaics and cases of tissue-specific mosaicism is suggested to result from trisomy rescue due to placental mosaicism [3, 4, 27, 31, 53], but it appears to be not the case of spontaneous abortions [50]. In neurodegenerative and aging diseases, somatic aneuploidy is probably the result of cellular natural selection — abnormal cells possess the potential to survive and to proliferate [8, 10, 45]. Studies of somatic cell division (mitosis) suggest that mitotic non-disjunction and anaphase lagging are two main mechanisms for post-zygotic aneuploidy formation (aneuploidization). Numerous intracellular processes are assumed to be involved in improper somatic cell divisions producing GIN and CIN. Among these are defects in kinetochore apparatus, centrosomes amplification, genetic and epigenetic alterations to mitotic checkpoint genes (aneuploidy/polyploidy) as well as abnormal DNA reparation and replication (structural alterations to chromosomes, aneuploidy, polyploidy) [3, 11, 45, 90-96]. Polyplodization followed by multipolar cellular divisions are also hypothesized to be a major contributor to somatic aneuploidization associated with human diseases [96]. Nonetheless, there is still a lack of an integrated view on SGV formation.

CONCLUDING REMARKS

In a previous issue of Current Genomics, we have hypothesized that uniqueness of a cell is achieved via SGV [3]. Single-cell gene expression studies showing that there is no an average cell, because each one has own unique epigenetic profiling (or epigenome) [21, 97]. Here, we would like repeatedly adopt this idea to the cellular genome. Four years after the first postulation [3], important additional data on SGV contribution to normal and pathological human biodiversity have been accumulated. It was found that early human prenatal development was defined as a major source for SGV [5, 6]. It has been shown that neurodegeneration is mediated by GIN and CIN like in cancer in such devastative genetic brain diseases as ataxia-telangiectasia and Alzheimer’s disease [10, 45]. Several psychiatric diseases (autism and schizophrenia) have been associated with mosaic (somatic) aneuploidy [35, 44]. Additionally, very recent reports provided by others groups of researchers showed SGV implicated in the normal and abnormal brain physiology and aging [98-100]. These results provide essential evidences that neuronal DNA variation is a new feature of the human brain, which may contribute to neural diversity in normal and pathophysiological states and differences amongst individuals. Together, one can conclude that SGV research has proven itself sufficiently to become an important biomedical field that would help to understand cellular and molecular processes determining human life- and health-span.

ACKNOWLEDGEMENTS

This article is dedicated to Dr. Ilia V. Soloviev. The authors are supported by Philip Morris USA Inc. and DLR/BMBF (RUS 09/006).

REFERENCES

  • 1.Los FJ, Van Opstal D, van den Berg C. The development of cytogenetically normal, abnormal and mosaic embryos: a theoretical model. Hum. Reprod. Update. 2004;10:79–94. doi: 10.1093/humupd/dmh005. [DOI] [PubMed] [Google Scholar]
  • 2.Erickson RP. Somatic gene mutation and human disease other than cancer: an update. Mutat. Res. 2010 doi: 10.1016/j.mrrev.2010.04.002. (in press) [DOI] [PubMed] [Google Scholar]
  • 3.Iourov IY, Vorsanova SG, Yurov YB. Intercellular gnomic (chromosomal) variations resulting in somatic mosaicism: mechanisms and consequences. Curr. Genomics. 2006;7:435–446. [Google Scholar]
  • 4.Iourov IY, Vorsanova SG, Yurov YB. Chromosomal variation in mammalian neuronal cells: known facts and attractive hypotheses. Int. Rev. Cytol. 2006;249:143–191. doi: 10.1016/S0074-7696(06)49003-3. [DOI] [PubMed] [Google Scholar]
  • 5.Yurov YB, Iourov IY, Vorsanova SG, Liehr T, Kolotii AD, Kutsev SI, Pellestor F, Beresheva AK, Demidova IA, Kravets VS, Monakhov VV, Soloviev IV. 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]
  • 6.Vanneste E, Voet T, Le Caignec C, Ampe M, Konings P, Melotte C, Debrock S, Amyere M, Vikkula M, Schuit F, Fryns JP, Verbeke G, D'Hooghe T, Moreau Y, Vermeesch JR. Chromosome instability is common in human cleavage-stage embryos. Nat. Med. 2009;15:577–583. doi: 10.1038/nm.1924. [DOI] [PubMed] [Google Scholar]
  • 7.Iourov IY, Vorsanova SG, Yurov YB. Chromosomal mosaicism goes global. Mol. Cytogenet. 2008;1:26. doi: 10.1186/1755-8166-1-26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Yurov YB, Vorsanova SG, Iourov IY. GIN'n'CIN hypothesis of brain aging: deciphering the role of somatic genetic instabilities and neural aneuploidy during ontogeny. Mol. Cytogenet. 2009;2:23. doi: 10.1186/1755-8166-2-23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Jackson SP, Bartek J. The DNA-damage response in human biology and disease. Nature. 2009;461:1071–1078. doi: 10.1038/nature08467. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.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;34:212–220. doi: 10.1016/j.nbd.2009.01.003. [DOI] [PubMed] [Google Scholar]
  • 11.Finkel T, Serrano M, Blasco MA. The common biology of cancer and ageing. Nature. 2007;448:767–774. doi: 10.1038/nature05985. [DOI] [PubMed] [Google Scholar]
  • 12.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:S186–S190. doi: 10.1016/s0387-7604(01)00363-1. [DOI] [PubMed] [Google Scholar]
  • 13.Youssoufian H, Pyeritz RE. Mechanisms and consequences of somatic mosaicism in humans. Nat. Rev. Genet. 2002;3:749–758. doi: 10.1038/nrg906. [DOI] [PubMed] [Google Scholar]
  • 14.Duesberg P. Chromosomal chaos and cancer. Sci. Am. 2007;296:52–59. doi: 10.1038/scientificamerican0507-52. [DOI] [PubMed] [Google Scholar]
  • 15.Iourov IY, Vorsanova SG, Yurov YB. Molecular cytogenetics and cytogenomics of brain diseases. Curr. Genomics. 2008;9:452–465. doi: 10.2174/138920208786241216. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Notini AJ, Craig JM, White SJ. Copy number variation and mosaicism. Cytogenet. Genome Res. 2008;123:270–277. doi: 10.1159/000184717. [DOI] [PubMed] [Google Scholar]
  • 17.Dierssen M, Herault Y, Estivill X. Aneuploidy: from a physiological mechanism of variance to Down syndrome. Physiol. Rev. 2009;89:887–920. doi: 10.1152/physrev.00032.2007. [DOI] [PubMed] [Google Scholar]
  • 18.Iourov IY, Vorsanova SG, Yurov YB, editors. Genomic landscape of the Alzheimer’s disease brain: chromosome instability — aneuploidy, but not tetraploidy — mediates neurodegeneration. Neurodegener. Dis. 2010 doi: 10.1159/000315398. (in press) [DOI] [PubMed] [Google Scholar]
  • 19.Gottlieb B, Beitel LK, Alvarado C, Trifiro MA. Selection and mutation in the "new" genetics: an emerging hypothesis. Hum. Genet. 2010;127:491–501. doi: 10.1007/s00439-010-0792-9. [DOI] [PubMed] [Google Scholar]
  • 20.Pleasance ED, Cheetham RK, Stephens PJ, McBride DJ, Humphray SJ, Greenman CD, Varela I, Lin ML, Ordóñez GR, Bignell GR, Ye K, Alipaz J, Bauer MJ, Beare D, Butler A, Carter RJ, Chen L, Cox AJ, Edkins S, Kokko-Gonzales PI, Gormley NA, Grocock RJ, Haudenschild CD, Hims MM, James T, Jia M, Kingsbury Z, Leroy C, Marshall J, Menzies A, Mudie LJ, Ning Z, Royce T, Schulz-Trieglaff OB, Spiridou A, Stebbings LA, Szajkowski L, Teague J, Williamson D, Chin L, Ross MT, Campbell PJ, Bentley DR, Futreal PA, Stratton MR. A comprehensive catalogue of somatic mutations from a human cancer genome. Nature. 2010;463:191–196. doi: 10.1038/nature08658. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Sgaramella V, Astolfi PA. Somatic genome variations interact with environment, genome and epigenome in the determination of the phenotype: a paradigm shift in genomics? DNA Repair. 2010;9:470–473. doi: 10.1016/j.dnarep.2009.11.011. [DOI] [PubMed] [Google Scholar]
  • 22.Vorsanova SG, Yurov YB, Iourov IY. Human interphase chromosomes: a review of available molecular cytogenetic technologies. Mol. Cytogenet. 2010;3:1. doi: 10.1186/1755-8166-3-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Redon R, Ishikawa S, Fitch KR, Feuk L, Perry GH, An-drews TD, Fiegler H, Shapero MH, Carson AR, Chen W, Cho EK, Dallaire S, Freeman JL, González JR, Gratacòs M, Huang J, Kalaitzopoulos D, Komura D, MacDonald JR, Marshall CR, Mei R, Montgomery L, Nishimura K, Okamura K, Shen F, Somerville MJ, Tchinda J, Valsesia A, Woodwark C, Yang F, Zhang J, Zerjal T, Zhang J, Armengol L, Conrad DF, Estivill X, Tyler-Smith C, Carter NP, Aburatani H, Lee C, Jones KW, Scherer SW, Hurles ME. Global variation in copy number in the human genome. Nature. 2006;444:444–454. doi: 10.1038/nature05329. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Scherer SW, Lee C, Birney E, Altshuler DM, Eichler EE, Carter NP, Hurles ME, Feuk L. Challenges and standards in integrating surveys of structural variation. Nat. Genet. 2007;39(7 Suppl):S7–S15. doi: 10.1038/ng2093. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Delhanty JD. Mechanisms of aneuploidy induction in human oogenesis and early embryogenesis. Cytogenet. Genome Res. 2005;111:237–244. doi: 10.1159/000086894. [DOI] [PubMed] [Google Scholar]
  • 26.Weier JF, Ferlatte C, Baumgartner A, Jung CJ, Nguyen HN, Chu LW, Pedersen RA, Fisher SJ, Weier HU. Molecular cytogenetic studies towards the full karyotype analysis of human blastocysts and cytotrophoblasts. Cytogenet. Genome Res. 2006;114:302–311. doi: 10.1159/000094218. [DOI] [PubMed] [Google Scholar]
  • 27.Weier JF, Weier HU, Jung CJ, Gormley M, Zhou Y, Chu LW, Genbacev O, Wright AA, Fisher SJ. Human cytotrophoblasts acquire aneuploidies as they differentiate to an invasive phenotype. Dev. Biol. 2005;279:420–432. doi: 10.1016/j.ydbio.2004.12.035. [DOI] [PubMed] [Google Scholar]
  • 28.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;53:385–390. doi: 10.1369/jhc.4A6430.2005. [DOI] [PubMed] [Google Scholar]
  • 29.Hultén MA, Patel SD, Westgren M, Papadogiannakis N, Jonsson AM, Jonasson J, Iwarsson E. On the paternal origin of trisomy 21 Down syndrome. Mol. Cytogenet. 2010;3:4. doi: 10.1186/1755-8166-3-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Hsu LY, Perlis TE. United States survey on chromosome mosaicism and pseudomosaicism in prenatal diagnosis. Prenat. Diagn. 1984;4(Spec No):97–130. doi: 10.1002/pd.1970040708. [DOI] [PubMed] [Google Scholar]
  • 31.Wolstenholme J. Confined placental mosaicism for trisomies 2, 3, 7, 8, 9, 16, and 22: their incidence, likely origins, and mechanisms for cell lineage compartmentalization. Prenat. Diagn. 1996;16:511–524. doi: 10.1002/(SICI)1097-0223(199606)16:6<511::AID-PD904>3.0.CO;2-8. [DOI] [PubMed] [Google Scholar]
  • 32.Stetten G, Escallon CS, South ST, McMichael JL, Saul DO, Blakemore KJ. Reevaluating confined placental mosaicism. Am. J. Med. Genet. A. 2004;131:232–239. doi: 10.1002/ajmg.a.30363. [DOI] [PubMed] [Google Scholar]
  • 33.Nielsen J, Wohlert M. Chromosome abnormalities found among 34,910 newborn children: results from a 13-year incidence study in Aarhus, Denmark. Hum. Genet. 1991;87:81–83. doi: 10.1007/BF01213097. [DOI] [PubMed] [Google Scholar]
  • 34.Jacobs PA, Browne C, Gregson N, Joyce C, White H. Estimates of the frequency of chromosome abnormalities detectable in unselected newborns using moderate levels of banding. J. Med. Genet. 1992;29:103–108. doi: 10.1136/jmg.29.2.103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Yurov YB, Vorsanova SG, Iourov IY, Demidova IA, Beresheva AK, Kravetz VS, Monakhov VV, Kolotii AD, Voinova-Ulas VY, Gorbachevskaya NL. Unexplained autism is frequently associated with low-level mosaic aneuploidy. J. Med. Genet. 2007;44:521–525. doi: 10.1136/jmg.2007.049312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Kleczkowska A, Fryns JP, van den Berghe H. On the variable effect of mosaic normal/balanced chromosomal rearrangements in man. J. Med. Genet. 1990;27:505–507. doi: 10.1136/jmg.27.8.505. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Nowinski GP, van Dyke DL, Tilley BC, Jacobsen G, Babu VR, Worsham MJ, Wilson GN, Weiss L. The frequency of aneuploidy in cultured lymphocytes is correlated with age and gender but not with reproductive history. Am. J. Hum. Genet. 1990;46:1101–1111. [PMC free article] [PubMed] [Google Scholar]
  • 38.Guttenbach M, Koschorz B, Bernthaler U, Grimm T, Schmid M. Sex chromosomes loss and aging: in situ hybridization studies on human interphase nuclei. Am. J. Hum. Genet. 1995;57:1143–1150. [PMC free article] [PubMed] [Google Scholar]
  • 39.Russel LM, Strike P, Browne CE, Jacobs PA. X chromosome loss and aging. Cytogenet. Genome. Res. 2007;116:181–185. doi: 10.1159/000098184. [DOI] [PubMed] [Google Scholar]
  • 40.Pressl S, Edwards A, Stephan G. The influence of age, sex and smoking habits on the background level of FISH-detected translocations. Mutat. Res. 1999;442:89 – 95. doi: 10.1016/s1383-5718(99)00067-4. [DOI] [PubMed] [Google Scholar]
  • 41.Geigl JB, Langer S, Barwisch S, Pfleghaar K, Lederer G, Speicher MR. Analysis of gene expression patterns and chromosomal changes associated with aging. Cancer Res. 2004;64:8850–8557. doi: 10.1158/0008-5472.CAN-04-2151. [DOI] [PubMed] [Google Scholar]
  • 42.Wilkens L, Flemming P, Gebel M, Bleck J, Terkamp C, Wingen L, Kreipe H, Schlegelberger B. Induction of aneuploidy by increasing chromosomal instability during dedifferentiation of hepatocellular carcinoma. Proc. Natl. Acad. Sci. USA. 2004;101:1309–1314. doi: 10.1073/pnas.0305817101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.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;14:223–229. doi: 10.1007/s10577-006-1037-6. [DOI] [PubMed] [Google Scholar]
  • 44.Yurov YB, Iourov IY, Vorsanova SG, Demidova IA, Kravets VS, Beresheva AK, Kolotii AD, Monakhov VV, Uranova NA, Vostrikov VM, Soloviev IV, Liehr T. The schizophrenia brain exhibits low-level aneuploidy involving chromosome 1. Schizophr. Res. 2008;98:139–147. doi: 10.1016/j.schres.2007.07.035. [DOI] [PubMed] [Google Scholar]
  • 45.Iourov IY, Vorsanova SG, Liehr T, Kolotii AD, Yurov YB. Increased chromosome instability dramatically disrupts neural genome integrity and mediates cerebellar degeneration in the ataxia-telangiectasia brain. Hum. Mol. Genet. 2009;18:2656–2669. doi: 10.1093/hmg/ddp207. [DOI] [PubMed] [Google Scholar]
  • 46.Piotrowski A, Bruder CE, Andersson R, de Ståhl TD, Menzel U, Sandgren J, Poplawski A, von Tell D, Crasto C, Bogdan A, Bartoszewski R, Bebok Z, Krzyzanowski M, Jankowski Z, Partridge EC, Komorowski J, Dumanski JP. Somatic mosaicism for copy number variation in differentiated human tissues. Hum. Mutat. 2008;29:1118–1124. doi: 10.1002/humu.20815. [DOI] [PubMed] [Google Scholar]
  • 47.Mkrtchyan H, Gross M, Hinreiner S, Polytiko A, Manvelyan M, Mrasek K, Kosyakova N, Ewers E, Nelle H, Liehr T, Volleth M, Weise A. Early embryonic chromosome instability results in stable mosaic pattern in human tissues. PLoS ONE. 2010;5:e9591. doi: 10.1371/journal.pone.0009591. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Iourov IY, Vorsanova SG, Yurov YB. Developmental neural chromosome instability as a possible cause of childhood brain cancers. Med. Hypotheses. 2009;72:615–616. doi: 10.1016/j.mehy.2008.12.003. [DOI] [PubMed] [Google Scholar]
  • 49.Menten B, Maas N, Thienpont B, Buysse K, Vandesompele J, Melotte C, de Ravel T, Van Vooren S, Balikova I, Backx L, Janssens S, De Paepe A, De Moor B, Moreau Y, Marynen P, Fryns JP, Mortier G, Devriendt K, Speleman F, Vermeesch JR. Emerging patterns of cryptic chromosomal imbalance in patients with idiopathic mental retardation and multiple congenital anomalies: a new series of 140 patients and review of published reports. J. Med. Genet. 2006;43:625–633. doi: 10.1136/jmg.2005.039453. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Vorsanova SG, Kolotii AD, Iourov IY, Monakhov VV, Kirillova EA, Soloviev IV, Yurov YB. Evidence for high frequency of chromosomal mosaicism in spontaneous abortions revealed by interphase FISH analysis. J. Histochem. Cytochem. 2005;53:375–380. doi: 10.1369/jhc.4A6424.2005. [DOI] [PubMed] [Google Scholar]
  • 51.Vorsanova SG, Iourov IY, Voinova-Ulas VY, Weise A, Monakhov VV, Kolotii AD, Soloviev IV, Novikov PV, Yurov YB, Liehr T. Partial monosomy 7q34-qter and 21pter-q22.13 due to cryptic unbalanced translocation t(7;21) but not monosomy of the whole chromosome 21: a case report plus review of the literature. Mol. Cytogenet. 2008;1:13. doi: 10.1186/1755-8166-1-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Weier JF, Weier HU, Nureddin A, Pedersen RA, Racowsky C. Aneuploidy involving chromosome 1 in failed-fertilized human oocytes is unrelated to maternal age. J. Assist. Reprod. Genet. 2005;22:285–293. doi: 10.1007/s10815-005-5999-7. [DOI] [PubMed] [Google Scholar]
  • 53.Liehr T. Cytogenetic contribution to uniparental disomy (UPD) Mol. Cytogenet. 2010;3:8. doi: 10.1186/1755-8166-3-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Woods CG, Bankier A, Curry J, Sheffield LJ, Slaney SF, Smith K, Voullaire L, Wellesley D. Asymmetry and skin pigmentary anomalies in chromosome mosaicism. J. Med. Genet. 1994;31:694–701. doi: 10.1136/jmg.31.9.694. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Hassold T, Hall H, Hunt P. The origin of human aneuploidy: where we have been, where we are going. Hum. Mol. Genet. 2007;16:R203–R208. doi: 10.1093/hmg/ddm243. [DOI] [PubMed] [Google Scholar]
  • 56.Liehr T, Claussen U, Starke H. Small supernumerary marker chromosomes (sSMC) in humans. Cytogenet. Genome Res. 2004;107:55–67. doi: 10.1159/000079572. [DOI] [PubMed] [Google Scholar]
  • 57.Fickelscher I, Starke H, Schulze E, Ernst G, Kosyakova N, Mkrtchyan H, MacDermont K, Sebire N, Liehr T. A further case with a small supernumerary marker chromosome (sSMC) derived from chromosome 1 — evidence for high variability in mosaicism in different tissues of sSMC carriers. Prenat. Diagn. 2007;27:783–785. doi: 10.1002/pd.1776. [DOI] [PubMed] [Google Scholar]
  • 58.Liehr T, Weise A. Frequency of small supernumerary marker chromosomes in prenatal, newborn, developmentally retarded and infertility diagnostics. Int. J. Mol. Med. 2007;19:719–731. [PubMed] [Google Scholar]
  • 59.Vorsanova SG, Iurov (Iourov) IY, Demidova IA, Voinova-Ulas VIu, Kravets VS, Solov'ev IV, Gorbachevskaia NL, Iurov (Yurov) YB. Variations of heterochromatic chromosomal regions and chromosome abnormalities in children with autism: identification of genetic markers in autistic spectrum disorders. Zh. Nevrol. Psikhiatr. Im. S S Korsakova. 2006;106:52–57. [PubMed] [Google Scholar]
  • 60.Vorsanova SG, Yurov IY, Demidova IA, Voinova-Ulas VY, Kravets VS, Solov'ev IV, Gorbachevskaya NL, Yurov YB. Variability in the heterochromatin regions of the chromosomes and chromosomal anomalies in children with autism: identification of genetic markers of autistic spectrum disorders. Neurosci. Behav. Physiol. 2007;37:553–558. doi: 10.1007/s11055-007-0052-1. [DOI] [PubMed] [Google Scholar]
  • 61.Vorsanova SG, Voinova-Ulas VIu, Iurov (Iourov) IY, Kurinnaia OS, Demidova IA, Iurov (Yurov) YB. Cytogenetic, molecular cytogenetic, clinical and genealogical study of mothers of children with autism: a search for family genetic markers of autistic disorders. Zh. Nevrol. Psikhiatr. Im. S S Korsakova. 2009;109:54–64. [PubMed] [Google Scholar]
  • 62.Carreira IM, Melo JB, Rodrigues C, Backx L, Vermeesch J, Weise A, Kosyakova N, Oliveira G, Matoso E. Molecular cytogenetic characterisation of a mosaic add(12)(p13.3) with an inv dup(3)(q26.31_>qter) detected in an autistic boy. Mol. Cytogenet. 2009;2:16. doi: 10.1186/1755-8166-2-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Castermans D, Thienpont B, Volders K, Crepel A, Vermeesch JR, Schrander-Stumpel CT, Van de Ven WJ, Steyaert JG, Creemers JW, Devriendt K. Position effect leading to haploinsufficiency in a mosaic ring chromosome 14 in a boy with autism. Eur. J. Hum. Genet. 2008;16:1187–1192. doi: 10.1038/ejhg.2008.71. [DOI] [PubMed] [Google Scholar]
  • 64.Wu DJ, Wang NJ, Driscoll J, Dorrani N, Liu D, Sigman M, Schanen NC. Autistic disorder associated with a paternally derived unbalanced translocation leading to duplication of chromosome 15pter-q13.2: a case report. Mol. Cytogenet. 2009;2:27. doi: 10.1186/1755-8166-2-27. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Hogart A, Wu D, LaSalle JM, Schanen NC. The comorbidity of autism with the genomic disorders of chromosome 15q11.2-q13. Neurobiol. Dis. 2010;38:181–191. doi: 10.1016/j.nbd.2008.08.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Havlovicova M, Novotna D, Kocarek E, Novotna K, Bendova S, Petrak B, Hrdlicka M, Sedlacek Z. A girl with neurofibromatosis type 1, atypical autism and mosaic ring chromosome 17. Am. J. Med. Genet. A. 2007;143:76–81. doi: 10.1002/ajmg.a.31569. [DOI] [PubMed] [Google Scholar]
  • 67.Fryns JP, Kleczkowska A. Autism and ring chromosome 18 mosaicism. Clin. Genet. 1992;42:55. doi: 10.1111/j.1399-0004.1992.tb03139.x. [DOI] [PubMed] [Google Scholar]
  • 68.Ghaziuddin M, Sheldon S, Tsai LY, Alessi N. Abnormalities of chromosome 18 in a girl with mental retardation and autistic disorder. J. Intellect. Disabil. Res. 1993;37:313–317. doi: 10.1111/j.1365-2788.1993.tb01288.x. [DOI] [PubMed] [Google Scholar]
  • 69.Sauter S, von Beust G, Burfeind P, Weise A, Starke H, Liehr T, Zoll B. Autistic disorder and chromosomal mosaicism 46,XY[123]/46,XY,del(20)(pter_>p12.2)[10] Am. J. Med. Genet. A. 2003;120A:533–536. doi: 10.1002/ajmg.a.20089. [DOI] [PubMed] [Google Scholar]
  • 70.Arrieta I, Nunez T, Martinez B, Perez A, Telez M, Criado B, Gainza I, Lostao CM. Chromsomal fragility in a behavioral disorder. Behav. Genet. 2002;32:397–412. doi: 10.1023/a:1020876010236. [DOI] [PubMed] [Google Scholar]
  • 71.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]
  • 72.DeLisi LE, Friedrich U, Wahlstrom J, Boccio-Smith A, Forsman A, Eklund K, Crow TJ. Schizophrenia and sex chromosome anomalies. Schizophr. Bull. 1994;20:495–505. doi: 10.1093/schbul/20.3.495. [DOI] [PubMed] [Google Scholar]
  • 73.Kumra S, Wiggs E, Krasnewich D, Meck J, Smith AC, Bedwell J, Fernandez T, Jacobsen LK, Lenane M, Rapoport JL. Brief report: association of sex chromosome anomalies with childhood-onset psychotic disorders. J. Am. Acad. Child Adolesc. Psychiatry. 1998;37:292–296. doi: 10.1097/00004583-199803000-00014. [DOI] [PubMed] [Google Scholar]
  • 74.Bassett AS, Chow ES, Weksberg R. Chromosomal abnormalites 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]
  • 75.Toyota T, Shimizu H, Yamada K, Yoshitsugu K, Meerbux J, Hattori E, Ichimiza T, Yoshikawa T. Karyotype analysis of 161 unrelated schizophrenics: no increased rates of chromosome X mosaicism, using ethnically matched and age-stratified controls. Schizophr. Res. 2001;52:171–179. doi: 10.1016/s0920-9964(01)00151-7. [DOI] [PubMed] [Google Scholar]
  • 76.Iurov (Iourov) IY, Vorsanova SG, Iurov (Yurov) YB. Chromosome abnormalities in schizophrenia. Zh. Nevrol. Psikhiatr. Im. S. S. Korsakova. 2006;106:75–82. [PubMed] [Google Scholar]
  • 77.Chen CH, Shih HH, Wang-Wuu S, Tai JJ, Wuu KD. Chromosomal fragile site expression in lymphocytes of schizophrenia. Hum. Genet. 1998;103:702–706. doi: 10.1007/s004390050894. [DOI] [PubMed] [Google Scholar]
  • 78.Beck JA, Poulter M, Campbell TA, Uphill JB, Adamson G, Geddes JF, Revesz T, Davis MB, Wood NW, Collinge J, Tabrizi SJ. Somatic and germline mosaicism in sporadic early-onset Alzheimer's disease. Hum. Mol. Genet. 2004;13:1219–24. doi: 10.1093/hmg/ddh134. [DOI] [PubMed] [Google Scholar]
  • 79.Swami M, Hendricks AE, Gillis T, Massood T, Mysore J, Myers RH, Wheeler VC. Somatic expansion of the Huntington's disease CAG repeat in the brain is associated with an earlier age of disease onset. Hum. Mol. Genet. 2009;18:3039–3047. doi: 10.1093/hmg/ddp242. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.De Biase I, Rasmussen A, Monticelli A, Al-Mahdawi S, Pook M, Cocozza S, Bidichandani SI. Somatic instability of the expanded GAA triplet-repeat sequence in Friedreich ataxia progresses throughout life. Genomics. 2007;90:1–5. doi: 10.1016/j.ygeno.2007.04.001. [DOI] [PubMed] [Google Scholar]
  • 81.Wada T, Candotti F. Somatic mosaicism in primary immune deficiencies. Curr. Opin. Allergy Clin. Immunol. 2008;8:510–514. doi: 10.1097/ACI.0b013e328314b651. [DOI] [PubMed] [Google Scholar]
  • 82.Invernizzi P, Miozzo M, Battezzati PM, Bianchi I, Grati FR, Simoni G, Selmi C, Watnik M, Gershwin ME, Podda M. Frequency of monosomy X in women with primary biliary cirrhosis. Lancet. 2004;363:533–535. doi: 10.1016/S0140-6736(04)15541-4. [DOI] [PubMed] [Google Scholar]
  • 83.Invernizzi P, Miozzo M, Selmi C, Persani L, Battezzati PM, Zuin M, Lucchi S, Meroni PL, Marasini B, Zeni S, Watnik M, Grati FR, Simoni G, Gershwin ME, Podda M. X chromosome monosomy: a common mechanism for autoimmune diseases. J. Immunol. 2005;175:575–578. doi: 10.4049/jimmunol.175.1.575. [DOI] [PubMed] [Google Scholar]
  • 84.Weismann CG, Gelb BD. The genetics of congenital heart disease: a review of recent developments. Curr. Opin. Cardiol. 2007;22:200–206. doi: 10.1097/HCO.0b013e3280f629c7. [DOI] [PubMed] [Google Scholar]
  • 85.Li L, McCormack AA, Nicholson JM, Fabarius A, Hehlmann R, Sachs RK, Duesberg PH. Cancer-causing karyotypes: chromosomal equilibria between destabilizing aneuploidy and stabilizing selection for oncogenic function. Cancer Genet. Cytogenet. 2009;188:1–25. doi: 10.1016/j.cancergencyto.2008.08.016. [DOI] [PubMed] [Google Scholar]
  • 86.Stankiewicz P, Lupski JR. Structural variation in the human genome and its role in disease. Annu. Rev. Med. 2010;61:437–455. doi: 10.1146/annurev-med-100708-204735. [DOI] [PubMed] [Google Scholar]
  • 87.Vorsanova SG, Iourov IY, Beresheva AK, Demidova IA, Monakhov VV, Kravets VS, Bartseva OB, Goyko EA, So-loviev IV, Yurov YB. Non-disjunction of chromosome 21, alphoid DNA variation, and sociogenetic features of Down syndrome. Tsitol. Genet. 2005;39(6):30–36. [PubMed] [Google Scholar]
  • 88.Iourov IY, Yurov YB, Vorsanova SG. Mosaic X chromosome aneuploidy can help to explain the male-to-female ratio in autism. Med. Hypotheses. 2008;70:456. doi: 10.1016/j.mehy.2007.05.037. [DOI] [PubMed] [Google Scholar]
  • 89.Vorsanova SG, Yurov YB, Iourov IY. Maternal smoking as a cause of mosaic aneuploidy in spontaneous abortions. Med. Hypotheses. 2008;71:607. doi: 10.1016/j.mehy.2008.05.027. [DOI] [PubMed] [Google Scholar]
  • 90.Cimini D, Degrassi F. Aneuploidy: a matter of bad connections. Trends Cell Biol. 2005;15:442–451. doi: 10.1016/j.tcb.2005.06.008. [DOI] [PubMed] [Google Scholar]
  • 91.Salmon ED, Cimini D, Cameron LA, DeLuca JG. Merotelic kinetochores in mammalian tissue cells. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2005;360:553–568. doi: 10.1098/rstb.2004.1610. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Weaver BA, Cleveland DW. Does aneuploidy cause cancer? Curr. Opin. Cell. Biol. 2006;18:658–667. doi: 10.1016/j.ceb.2006.10.002. [DOI] [PubMed] [Google Scholar]
  • 93.Herrera LA, Prada D, Andonegui MA, Duenas-Gonzales A. The epigenetic origin of aneuploidy. Curr. Genomics. 2008;9:43–50. doi: 10.2174/138920208783884883. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Chanhok NS, Pellman D. A little CIN may cost a lot: revisiting aneuploidy and cancer. Curr. Opin. Genet. Dev. 2009;19:74–81. doi: 10.1016/j.gde.2008.12.004. [DOI] [PubMed] [Google Scholar]
  • 95.Schvarzman JM, Sotillo R, Benezra R. Mitotic chromosomal instability and cancer: mouse modeling of the human disease. Nat. Rev. Cancer. 2010;10:102–115. doi: 10.1038/nrc2781. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Storchova Z, Pellman D. From polyploidy to aneuploidy, genome instability and cancer. Nat. Rev. Mol. Cell Biol. 2004;5:45–54. doi: 10.1038/nrm1276. [DOI] [PubMed] [Google Scholar]
  • 97.Levsky JM, Singer RH. Gene expression and the myth of the average cell. Trends Cell Biol. 2003;13:4–6. doi: 10.1016/s0962-8924(02)00002-8. [DOI] [PubMed] [Google Scholar]
  • 98.Arendt T, Mosch B, Morawski M. Neuronal aneuploidy in health and disease: cytomic approach to understand the molecular individuality of neurons. Int. J. Mol. Sci. 2009;10(4):1609–1627. doi: 10.3390/ijms10041609. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Arendt T, Brückner MK, Mosch B, Lösche A. Selective Cell Death of Hyperploid Neurons in Alzheimer's Disease. Am. J. Pathol. 2010 doi: 10.2353/ajpath.2010.090955. (in press) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Westra JW, Rivera RR, Bushman DM, Yung YC, Peterson SE, Barral S, Chun J. Neuronal DNA content variation (DCV) with regional and individual differences in the human brain. J. Comp. Neurol. 2010 doi: 10.1002/cne.22436. (in press.) [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Current Genomics are provided here courtesy of Bentham Science Publishers

RESOURCES