fbpx
Wikipedia

CpG site

The CpG sites or CG sites are regions of DNA where a cytosine nucleotide is followed by a guanine nucleotide in the linear sequence of bases along its 5' → 3' direction. CpG sites occur with high frequency in genomic regions called CpG islands (or CG islands).

a CpG site, i.e., the " 5'—C—phosphate—G—3' " sequence of nucleotides, is indicated on one DNA strand (in yellow). On the reverse DNA strand (in blue), the complementary 5'—CpG—3' site is shown. A C-G base-pairing between the two DNA strands is also indicated (right)

Cytosines in CpG dinucleotides can be methylated to form 5-methylcytosines. Enzymes that add a methyl group are called DNA methyltransferases. In mammals, 70% to 80% of CpG cytosines are methylated.[1] Methylating the cytosine within a gene can change its expression, a mechanism that is part of a larger field of science studying gene regulation that is called epigenetics. Methylated cytosines often mutate to thymines.

In humans, about 70% of promoters located near the transcription start site of a gene (proximal promoters) contain a CpG island.[2][3]

CpG characteristics

Definition

CpG is shorthand for 5'—C—phosphate—G—3' , that is, cytosine and guanine separated by only one phosphate group; phosphate links any two nucleosides together in DNA. The CpG notation is used to distinguish this single-stranded linear sequence from the CG base-pairing of cytosine and guanine for double-stranded sequences. The CpG notation is therefore to be interpreted as the cytosine being 5 prime to the guanine base. CpG should not be confused with GpC, the latter meaning that a guanine is followed by a cytosine in the 5' → 3' direction of a single-stranded sequence.

Under-representation caused by high mutation rate

CpG dinucleotides have long been observed to occur with a much lower frequency in the sequence of vertebrate genomes than would be expected due to random chance. For example, in the human genome, which has a 42% GC content,[4] a pair of nucleotides consisting of cytosine followed by guanine would be expected to occur   of the time. The frequency of CpG dinucleotides in human genomes is less than one-fifth of the expected frequency.[5]

This underrepresentation is a consequence of the high mutation rate of methylated CpG sites: the spontaneously occurring deamination of a methylated cytosine results in a thymine, and the resulting G:T mismatched bases are often improperly resolved to A:T; whereas the deamination of unmethylated cytosine results in a uracil, which as a foreign base is quickly replaced by a cytosine by the base excision repair mechanism. The C to T transition rate at methylated CpG sites is ~10 fold higher than at unmethylated sites.[6][7][8][9]

Genomic distribution

CpG sites GpC sites
   
Distribution of CpG sites (left: in red) and GpC sites (right: in green) in the human APRT gene. CpG are more abundant in the upstream region of the gene, where they form a CpG island, whereas GpC are more evenly distributed. The 5 exons of the APRT gene are indicated (blue), and the start (ATG) and stop (TGA) codons are emphasized (bold blue).

CpG dinucleotides frequently occur in CpG islands (see definition of CpG islands, below). There are 28,890 CpG islands in the human genome, (50,267 if one includes CpG islands in repeat sequences).[10] This is in agreement with the 28,519 CpG islands found by Venter et al.[11] since the Venter et al. genome sequence did not include the interiors of highly similar repetitive elements and the extremely dense repeat regions near the centromeres.[12] Since CpG islands contain multiple CpG dinucleotide sequences, there appear to be more than 20 million CpG dinucleotides in the human genome.

CpG islands

 
How methylation of CpG sites followed by spontaneous deamination leads to a lack of CpG sites in methylated DNA. As a result, residual CpG islands are created in areas where methylation is rare, and CpG sites stick (or where C to T mutation is highly detrimental).

CpG islands (or CG islands) are regions with a high frequency of CpG sites. Though objective definitions for CpG islands are limited, the usual formal definition is a region with at least 200 bp, a GC percentage greater than 50%, and an observed-to-expected CpG ratio greater than 60%. The "observed-to-expected CpG ratio" can be derived where the observed is calculated as:   and the expected as  [13] or  .[14]

Many genes in mammalian genomes have CpG islands associated with the start of the gene[15] (promoter regions). Because of this, the presence of a CpG island is used to help in the prediction and annotation of genes.

In mammalian genomes, CpG islands are typically 300–3,000 base pairs in length, and have been found in or near approximately 40% of promoters of mammalian genes.[16] Over 60% of human genes and almost all house-keeping genes have their promoters embedded in CpG islands.[17] Given the frequency of GC two-nucleotide sequences, the number of CpG dinucleotides is much lower than would be expected.[14]

A 2002 study revised the rules of CpG island prediction to exclude other GC-rich genomic sequences such as Alu repeats. Based on an extensive search on the complete sequences of human chromosomes 21 and 22, DNA regions greater than 500 bp were found more likely to be the "true" CpG islands associated with the 5' regions of genes if they had a GC content greater than 55%, and an observed-to-expected CpG ratio of 65%.[18]

CpG islands are characterized by CpG dinucleotide content of at least 60% of that which would be statistically expected (~4–6%), whereas the rest of the genome has much lower CpG frequency (~1%), a phenomenon called CG suppression. Unlike CpG sites in the coding region of a gene, in most instances the CpG sites in the CpG islands of promoters are unmethylated if the genes are expressed. This observation led to the speculation that methylation of CpG sites in the promoter of a gene may inhibit gene expression. Methylation, along with histone modification, is central to imprinting.[19] Most of the methylation differences between tissues, or between normal and cancer samples, occur a short distance from the CpG islands (at "CpG island shores") rather than in the islands themselves.[20]

CpG islands typically occur at or near the transcription start site of genes, particularly housekeeping genes, in vertebrates.[14] A C (cytosine) base followed immediately by a G (guanine) base (a CpG) is rare in vertebrate DNA because the cytosines in such an arrangement tend to be methylated. This methylation helps distinguish the newly synthesized DNA strand from the parent strand, which aids in the final stages of DNA proofreading after duplication. However, over time methylated cytosines tend to turn into thymines because of spontaneous deamination. There is a special enzyme in humans (Thymine-DNA glycosylase, or TDG) that specifically replaces T's from T/G mismatches. However, due to the rarity of CpGs, it is theorised to be insufficiently effective in preventing a possibly rapid mutation of the dinucleotides. The existence of CpG islands is usually explained by the existence of selective forces for relatively high CpG content, or low levels of methylation in that genomic area, perhaps having to do with the regulation of gene expression. A 2011 study showed that most CpG islands are a result of non-selective forces.[21]

Methylation, silencing, cancer, and aging

 
An image showing a hypothetical evolutionary mechanism behind CpG island formation.

CpG islands in promoters

In humans, about 70% of promoters located near the transcription start site of a gene (proximal promoters) contain a CpG island.[2][3]

Distal promoter elements also frequently contain CpG islands. An example is the DNA repair gene ERCC1, where the CpG island-containing element is located about 5,400 nucleotides upstream of the transcription start site of the ERCC1 gene.[22] CpG islands also occur frequently in promoters for functional noncoding RNAs such as microRNAs.[23]

Methylation of CpG islands stably silences genes

In humans, DNA methylation occurs at the 5 position of the pyrimidine ring of the cytosine residues within CpG sites to form 5-methylcytosines. The presence of multiple methylated CpG sites in CpG islands of promoters causes stable silencing of genes.[24] Silencing of a gene may be initiated by other mechanisms, but this is often followed by methylation of CpG sites in the promoter CpG island to cause the stable silencing of the gene.[24]

Promoter CpG hyper/hypo-methylation in cancer

In cancers, loss of expression of genes occurs about 10 times more frequently by hypermethylation of promoter CpG islands than by mutations. For example, in a colorectal cancer there are usually about 3 to 6 driver mutations and 33 to 66 hitchhiker or passenger mutations.[25] In contrast, in one study of colon tumors compared to adjacent normal-appearing colonic mucosa, 1,734 CpG islands were heavily methylated in tumors whereas these CpG islands were not methylated in the adjacent mucosa.[26] Half of the CpG islands were in promoters of annotated protein coding genes,[26] suggesting that about 867 genes in a colon tumor have lost expression due to CpG island methylation. A separate study found an average of 1,549 differentially methylated regions (hypermethylated or hypomethylated) in the genomes of six colon cancers (compared to adjacent mucosa), of which 629 were in known promoter regions of genes.[27] A third study found more than 2,000 genes differentially methylated between colon cancers and adjacent mucosa. Using gene set enrichment analysis, 569 out of 938 gene sets were hypermethylated and 369 were hypomethylated in cancers.[28] Hypomethylation of CpG islands in promoters results in overexpression of the genes or gene sets affected.

One 2012 study[29] listed 147 specific genes with colon cancer-associated hypermethylated promoters, along with the frequency with which these hypermethylations were found in colon cancers. At least 10 of those genes had hypermethylated promoters in nearly 100% of colon cancers. They also indicated 11 microRNAs whose promoters were hypermethylated in colon cancers at frequencies between 50% and 100% of cancers. MicroRNAs (miRNAs) are small endogenous RNAs that pair with sequences in messenger RNAs to direct post-transcriptional repression. On average, each microRNA represses several hundred target genes.[30] Thus microRNAs with hypermethylated promoters may be allowing over-expression of hundreds to thousands of genes in a cancer.

The information above shows that, in cancers, promoter CpG hyper/hypo-methylation of genes and of microRNAs causes loss of expression (or sometimes increased expression) of far more genes than does mutation.

DNA repair genes with hyper/hypo-methylated promoters in cancers

DNA repair genes are frequently repressed in cancers due to hypermethylation of CpG islands within their promoters. In head and neck squamous cell carcinomas at least 15 DNA repair genes have frequently hypermethylated promoters; these genes are XRCC1, MLH3, PMS1, RAD51B, XRCC3, RAD54B, BRCA1, SHFM1, GEN1, FANCE, FAAP20, SPRTN, SETMAR, HUS1, and PER1.[31] About seventeen types of cancer are frequently deficient in one or more DNA repair genes due to hypermethylation of their promoters.[32] As an example, promoter hypermethylation of the DNA repair gene MGMT occurs in 93% of bladder cancers, 88% of stomach cancers, 74% of thyroid cancers, 40%-90% of colorectal cancers and 50% of brain cancers. Promoter hypermethylation of LIG4 occurs in 82% of colorectal cancers. Promoter hypermethylation of NEIL1 occurs in 62% of head and neck cancers and in 42% of non-small-cell lung cancers. Promoter hypermethylation of ATM occurs in 47% of non-small-cell lung cancers. Promoter hypermethylation of MLH1 occurs in 48% of non-small-cell lung cancer squamous cell carcinomas. Promoter hypermethylation of FANCB occurs in 46% of head and neck cancers.

On the other hand, the promoters of two genes, PARP1 and FEN1, were hypomethylated and these genes were over-expressed in numerous cancers. PARP1 and FEN1 are essential genes in the error-prone and mutagenic DNA repair pathway microhomology-mediated end joining. If this pathway is over-expressed the excess mutations it causes can lead to cancer. PARP1 is over-expressed in tyrosine kinase-activated leukemias,[33] in neuroblastoma,[34] in testicular and other germ cell tumors,[35] and in Ewing's sarcoma,[36] FEN1 is over-expressed in the majority of cancers of the breast,[37] prostate,[38] stomach,[39][40] neuroblastomas,[41] pancreatic,[42] and lung.[43]

DNA damage appears to be the primary underlying cause of cancer.[44][45] If accurate DNA repair is deficient, DNA damages tend to accumulate. Such excess DNA damage can increase mutational errors during DNA replication due to error-prone translesion synthesis. Excess DNA damage can also increase epigenetic alterations due to errors during DNA repair.[46][47] Such mutations and epigenetic alterations can give rise to cancer (see malignant neoplasms). Thus, CpG island hyper/hypo-methylation in the promoters of DNA repair genes are likely central to progression to cancer.

Methylation of CpG sites with age

Since age has a strong effect on DNA methylation levels on tens of thousands of CpG sites, one can define a highly accurate biological clock (referred to as epigenetic clock or DNA methylation age) in humans and chimpanzees.[48]

Unmethylated sites

Unmethylated CpG dinucleotide sites can be detected by Toll-like receptor 9 (TLR 9)[49] on plasmacytoid dendritic cells, monocytes, natural killer (NK) cells, and B cells in humans. This is used to detect intracellular viral infection.

Role of CpG sites in memory

In mammals, DNA methyltransferases (which add methyl groups to DNA bases) exhibit a sequence preference for cytosines within CpG sites.[50] In the mouse brain, 4.2% of all cytosines are methylated, primarily in the context of CpG sites, forming 5mCpG.[51] Most hypermethylated 5mCpG sites increase the repression of associated genes.[51]

As reviewed by Duke et al., neuron DNA methylation (repressing expression of particular genes) is altered by neuronal activity. Neuron DNA methylation is required for synaptic plasticity; is modified by experiences; and active DNA methylation and demethylation is required for memory formation and maintenance.[52]

In 2016 Halder et al.[53] using mice, and in 2017 Duke et al.[52] using rats, subjected the rodents to contextual fear conditioning, causing an especially strong long-term memory to form. At 24 hours after the conditioning, in the hippocampus brain region of rats, the expression of 1,048 genes was down-regulated (usually associated with 5mCpG in gene promoters) and the expression of 564 genes was up-regulated (often associated with hypomethylation of CpG sites in gene promoters). At 24 hours after training, 9.2% of the genes in the rat genome of hippocampus neurons were differentially methylated. However while the hippocampus is essential for learning new information it does not store information itself. In the mouse experiments of Halder, 1,206 differentially methylated genes were seen in the hippocampus one hour after contextual fear conditioning but these altered methylations were reversed and not seen after four weeks. In contrast with the absence of long-term CpG methylation changes in the hippocampus, substantial differential CpG methylation could be detected in cortical neurons during memory maintenance. There were 1,223 differentially methylated genes in the anterior cingulate cortex of mice four weeks after contextual fear conditioning.

Demethylation at CpG sites requires ROS activity

 
Initiation of DNA demethylation at a CpG site.

In adult somatic cells DNA methylation typically occurs in the context of CpG dinucleotides (CpG sites), forming 5-methylcytosine-pG, or 5mCpG. Reactive oxygen species (ROS) may attack guanine at the dinucleotide site, forming 8-hydroxy-2'-deoxyguanosine (8-OHdG), and resulting in a 5mCp-8-OHdG dinucleotide site. The base excision repair enzyme OGG1 targets 8-OHdG and binds to the lesion without immediate excision. OGG1, present at a 5mCp-8-OHdG site recruits TET1 and TET1 oxidizes the 5mC adjacent to the 8-OHdG. This initiates demethylation of 5mC.[54]

 
Demethylation of 5-Methylcytosine (5mC) in neuron DNA.

As reviewed in 2018,[55] in brain neurons, 5mC is oxidized by the ten-eleven translocation (TET) family of dioxygenases (TET1, TET2, TET3) to generate 5-hydroxymethylcytosine (5hmC). In successive steps TET enzymes further hydroxylate 5hmC to generate 5-formylcytosine (5fC) and 5-carboxylcytosine (5caC). Thymine-DNA glycosylase (TDG) recognizes the intermediate bases 5fC and 5caC and excises the glycosidic bond resulting in an apyrimidinic site (AP site). In an alternative oxidative deamination pathway, 5hmC can be oxidatively deaminated by activity-induced cytidine deaminase/apolipoprotein B mRNA editing complex (AID/APOBEC) deaminases to form 5-hydroxymethyluracil (5hmU) or 5mC can be converted to thymine (Thy). 5hmU can be cleaved by TDG, single-strand-selective monofunctional uracil-DNA glycosylase 1 (SMUG1), Nei-Like DNA Glycosylase 1 (NEIL1), or methyl-CpG binding protein 4 (MBD4). AP sites and T:G mismatches are then repaired by base excision repair (BER) enzymes to yield cytosine (Cyt).

Two reviews[56][57] summarize the large body of evidence for the critical and essential role of ROS in memory formation. The DNA demethylation of thousands of CpG sites during memory formation depends on initiation by ROS. In 2016, Zhou et al.,[54] showed that ROS have a central role in DNA demethylation.

TET1 is a key enzyme involved in demethylating 5mCpG. However, TET1 is only able to act on 5mCpG if an ROS has first acted on the guanine to form 8-hydroxy-2'-deoxyguanosine (8-OHdG), resulting in a 5mCp-8-OHdG dinucleotide (see first figure in this section).[54] After formation of 5mCp-8-OHdG, the base excision repair enzyme OGG1 binds to the 8-OHdG lesion without immediate excision. Adherence of OGG1 to the 5mCp-8-OHdG site recruits TET1, allowing TET1 to oxidize the 5mC adjacent to 8-OHdG, as shown in the first figure in this section. This initiates the demethylation pathway shown in the second figure in this section.

Altered protein expression in neurons, controlled by ROS-dependent demethylation of CpG sites in gene promoters within neuron DNA, is central to memory formation.[58]

CpG loss

CpG depletion has been observed in the process of DNA methylation of Transposable Elements (TEs) where TEs are not only responsible in the genome expansion but also CpG loss in a host DNA. TEs can be known as "methylation centers" whereby the methylation process, the TEs spreads into the flanking DNA once in the host DNA. This spreading might subsequently result in CpG loss over evolutionary time. Older evolutionary times show a higher CpG loss in the flanking DNA, compared to the younger evolutionary times. Therefore, the DNA methylation can lead eventually to the noticeably loss of CpG sites in neighboring DNA. [59]

Genome size and CpG ratio are negatively correlated

 
CpG methylation contributes to the genome expansion and consequently to CpG depletion. This picture shows a genome with no TEs and unmethylated CpG sites, and the insertion and transposition of a TE lead to methylation and silencing of the TE. Through the process of CpG methylation a decrease in CpG is found.[60]

Previous studies have confirmed the variety of genomes sizes amount species, where invertebrates and vertebrates have small and big genomes compared to humans. The genome size is strongly connected to the number of transposable elements. However, there is a correlation between the number of TEs methylation versus the CpG amount. This negative correlation consequently causes depletion of CpG due to intergenic DNA methylation which is mostly attributed to the methylation of TEs. Overall, this contributes to a noticeable amount of CpG loss in different genomes species. [59]

Alu elements as promoters of CpG loss

Alu elements are known as the most abundant type of transposable elements. Some studies have used Alu elements as a way to study the idea of which factor is responsible for genome expansion. Alu elements are CpG-rich in a longer amount of sequence, unlike LINEs and ERVs. Alus can work as a methylation center, and the insertion into a host DNA can produce DNA methylation and provoke a spreading into the Flanking DNA area. This spreading is why there are a considerable amount CpG loss and a considerable increase in genome expansion.[59] However, this is a result that is analyzed over time because older Alus elements show more CpG loss in sites of neighboring DNA compared to younger ones.

See also

References

  1. ^ Jabbari K, Bernardi G (May 2004). "Cytosine methylation and CpG, TpG (CpA) and TpA frequencies". Gene. 333: 143–9. doi:10.1016/j.gene.2004.02.043. PMID 15177689.
  2. ^ a b Saxonov S, Berg P, Brutlag DL (2006). "A genome-wide analysis of CpG dinucleotides in the human genome distinguishes two distinct classes of promoters". Proc. Natl. Acad. Sci. U.S.A. 103 (5): 1412–7. Bibcode:2006PNAS..103.1412S. doi:10.1073/pnas.0510310103. PMC 1345710. PMID 16432200.
  3. ^ a b Deaton AM, Bird A (2011). "CpG islands and the regulation of transcription". Genes Dev. 25 (10): 1010–22. doi:10.1101/gad.2037511. PMC 3093116. PMID 21576262.
  4. ^ Lander, Eric S.; Linton, Lauren M.; Birren, Bruce; Nusbaum, Chad; Zody, Michael C.; Baldwin, Jennifer; Devon, Keri; Dewar, Ken; Doyle, Michael (15 February 2001). "Initial sequencing and analysis of the human genome". Nature. 409 (6822): 860–921. Bibcode:2001Natur.409..860L. doi:10.1038/35057062. ISSN 1476-4687. PMID 11237011.
  5. ^ International Human Genome Sequencing Consortium (2001-02-15). "Initial sequencing and analysis of the human genome". Nature. 409 (6822): 860–921. Bibcode:2001Natur.409..860L. doi:10.1038/35057062. ISSN 0028-0836. PMID 11237011.
  6. ^ Hwang DG, Green P (2004). "Bayesian Markov chain Monte Carlo sequence analysis reveals varying neutral substitution patterns in mammalian evolution". Proc Natl Acad Sci U S A. 101 (39): 13994–4001. Bibcode:2004PNAS..10113994H. doi:10.1073/pnas.0404142101. PMC 521089. PMID 15292512.
  7. ^ Walsh CP, Xu GL (2006). "Cytosine methylation and DNA repair". Curr Top Microbiol Immunol. Current Topics in Microbiology and Immunology. 301: 283–315. doi:10.1007/3-540-31390-7_11. ISBN 3-540-29114-8. PMID 16570853.
  8. ^ Arnheim N, Calabrese P (2009). "Understanding what determines the frequency and pattern of human germline mutations". Nat Rev Genet. 10 (7): 478–488. doi:10.1038/nrg2529. PMC 2744436. PMID 19488047.
  9. ^ Ségurel L, Wyman MJ, Przeworski M (2014). "Understanding what determines the frequency and pattern of human germline mutations". Annu Rev Genom Hum Genet. 15: 47–70. doi:10.1146/annurev-genom-031714-125740. PMID 25000986.
  10. ^ Lander ES, Linton LM, Birren B, Nusbaum C, Zody MC, Baldwin J, et al. (February 2001). "Initial sequencing and analysis of the human genome". Nature. 409 (6822): 860–921. Bibcode:2001Natur.409..860L. doi:10.1038/35057062. PMID 11237011.
  11. ^ Venter JC, Adams MD, Myers EW, Li PW, Mural RJ, Sutton GG, et al. (February 2001). "The sequence of the human genome". Science. 291 (5507): 1304–51. Bibcode:2001Sci...291.1304V. doi:10.1126/science.1058040. PMID 11181995.
  12. ^ Myers EW, Sutton GG, Smith HO, Adams MD, Venter JC (April 2002). "On the sequencing and assembly of the human genome". Proc. Natl. Acad. Sci. U.S.A. 99 (7): 4145–6. Bibcode:2002PNAS...99.4145M. doi:10.1073/pnas.092136699. PMC 123615. PMID 11904395.
  13. ^ Gardiner-Garden M, Frommer M (1987). "CpG islands in vertebrate genomes". Journal of Molecular Biology. 196 (2): 261–282. doi:10.1016/0022-2836(87)90689-9. PMID 3656447.
  14. ^ a b c Saxonov S, Berg P, Brutlag DL (2006). "A genome-wide analysis of CpG dinucleotides in the human genome distinguishes two distinct classes of promoters". Proc Natl Acad Sci USA. 103 (5): 1412–1417. Bibcode:2006PNAS..103.1412S. doi:10.1073/pnas.0510310103. PMC 1345710. PMID 16432200.
  15. ^ Hartl DL, Jones EW (2005). Genetics: Analysis of Genes and Genomes (6th ed.). Missisauga: Jones & Bartlett, Canada. p. 477. ISBN 978-0-7637-1511-3.
  16. ^ Fatemi M, Pao MM, Jeong S, Gal-Yam EN, Egger G, Weisenberger DJ, et al. (2005). "Footprinting of mammalian promoters: use of a CpG DNA methyltransferase revealing nucleosome positions at a single molecule level". Nucleic Acids Res. 33 (20): e176. doi:10.1093/nar/gni180. PMC 1292996. PMID 16314307.
  17. ^ Alberts, Bruce (18 November 2014). Molecular biology of the cell (Sixth ed.). New York, NY. p. 406. ISBN 978-0-8153-4432-2. OCLC 887605755.{{cite book}}: CS1 maint: location missing publisher (link)
  18. ^ Takai D, Jones PA (2002). "Comprehensive analysis of CpG islands in human chromosomes 21 and 22". Proc Natl Acad Sci USA. 99 (6): 3740–5. Bibcode:2002PNAS...99.3740T. doi:10.1073/pnas.052410099. PMC 122594. PMID 11891299.
  19. ^ Feil R, Berger F (2007). "Convergent evolution of genomic imprinting in plants and mammals". Trends Genet. 23 (4): 192–199. doi:10.1016/j.tig.2007.02.004. PMID 17316885.
  20. ^ Irizarry RA, Ladd-Acosta C, Wen B, Wu Z, Montano C, Onyango P, et al. (2009). "The human colon cancer methylome shows similar hypo- and hypermethylation at conserved tissue-specific CpG island shores". Nature Genetics. 41 (2): 178–186. doi:10.1038/ng.298. PMC 2729128. PMID 19151715.
  21. ^ Cohen N, Kenigsberg E, Tanay A (2011). "Primate CpG Islands Are Maintained by Heterogeneous Evolutionary Regimes Involving Minimal Selection". Cell. 145 (5): 773–786. doi:10.1016/j.cell.2011.04.024. PMID 21620139. S2CID 14856605.
  22. ^ Chen HY, Shao CJ, Chen FR, Kwan AL, Chen ZP (2010). "Role of ERCC1 promoter hypermethylation in drug resistance to cisplatin in human gliomas". Int. J. Cancer. 126 (8): 1944–54. doi:10.1002/ijc.24772. PMID 19626585.
  23. ^ Kaur S, Lotsari-Salomaa JE, Seppänen-Kaijansinkko R, Peltomäki P (2016). "MicroRNA Methylation in Colorectal Cancer". Adv. Exp. Med. Biol. Advances in Experimental Medicine and Biology. 937: 109–22. doi:10.1007/978-3-319-42059-2_6. ISBN 978-3-319-42057-8. PMID 27573897.
  24. ^ a b Bird A (2002). "DNA methylation patterns and epigenetic memory". Genes Dev. 16 (1): 6–21. doi:10.1101/gad.947102. PMID 11782440.
  25. ^ Vogelstein B, Papadopoulos N, Velculescu VE, Zhou S, Diaz LA, Kinzler KW (2013). "Cancer genome landscapes". Science. 339 (6127): 1546–58. Bibcode:2013Sci...339.1546V. doi:10.1126/science.1235122. PMC 3749880. PMID 23539594.
  26. ^ a b Illingworth RS, Gruenewald-Schneider U, Webb S, Kerr AR, James KD, Turner DJ, Smith C, Harrison DJ, Andrews R, Bird AP (2010). "Orphan CpG islands identify numerous conserved promoters in the mammalian genome". PLOS Genet. 6 (9): e1001134. doi:10.1371/journal.pgen.1001134. PMC 2944787. PMID 20885785.
  27. ^ Wei J, Li G, Dang S, Zhou Y, Zeng K, Liu M (2016). "Discovery and Validation of Hypermethylated Markers for Colorectal Cancer". Dis. Markers. 2016: 1–7. doi:10.1155/2016/2192853. PMC 4963574. PMID 27493446.
  28. ^ Beggs AD, Jones A, El-Bahrawy M, El-Bahwary M, Abulafi M, Hodgson SV, et al. (2013). "Whole-genome methylation analysis of benign and malignant colorectal tumours". J. Pathol. 229 (5): 697–704. doi:10.1002/path.4132. PMC 3619233. PMID 23096130.
  29. ^ Schnekenburger M, Diederich M (2012). "Epigenetics Offer New Horizons for Colorectal Cancer Prevention". Curr Colorectal Cancer Rep. 8 (1): 66–81. doi:10.1007/s11888-011-0116-z. PMC 3277709. PMID 22389639.
  30. ^ Friedman RC, Farh KK, Burge CB, Bartel DP (2009). "Most mammalian mRNAs are conserved targets of microRNAs". Genome Res. 19 (1): 92–105. doi:10.1101/gr.082701.108. PMC 2612969. PMID 18955434.
  31. ^ Rieke DT, Ochsenreither S, Klinghammer K, Seiwert TY, Klauschen F, Tinhofer I, et al. (2016). "Methylation of RAD51B, XRCC3 and other homologous recombination genes is associated with expression of immune checkpoints and an inflammatory signature in squamous cell carcinoma of the head and neck, lung and cervix". Oncotarget. 7 (46): 75379–75393. doi:10.18632/oncotarget.12211. PMC 5342748. PMID 27683114.
  32. ^ Jin B, Robertson KD (2013). "DNA methyltransferases, DNA damage repair, and cancer". Adv. Exp. Med. Biol. Advances in Experimental Medicine and Biology. 754: 3–29. doi:10.1007/978-1-4419-9967-2_1. ISBN 978-1-4419-9966-5. PMC 3707278. PMID 22956494.
  33. ^ Muvarak N, Kelley S, Robert C, Baer MR, Perrotti D, Gambacorti-Passerini C, et al. (2015). "c-MYC Generates Repair Errors via Increased Transcription of Alternative-NHEJ Factors, LIG3 and PARP1, in Tyrosine Kinase-Activated Leukemias". Mol. Cancer Res. 13 (4): 699–712. doi:10.1158/1541-7786.MCR-14-0422. PMC 4398615. PMID 25828893.
  34. ^ Newman EA, Lu F, Bashllari D, Wang L, Opipari AW, Castle VP (2015). "Alternative NHEJ Pathway Components Are Therapeutic Targets in High-Risk Neuroblastoma". Mol. Cancer Res. 13 (3): 470–82. doi:10.1158/1541-7786.MCR-14-0337. PMID 25563294.
  35. ^ Mego M, Cierna Z, Svetlovska D, Macak D, Machalekova K, Miskovska V, et al. (2013). "PARP expression in germ cell tumours". J. Clin. Pathol. 66 (7): 607–12. doi:10.1136/jclinpath-2012-201088. PMID 23486608. S2CID 535704.
  36. ^ Newman RE, Soldatenkov VA, Dritschilo A, Notario V (2002). "Poly(ADP-ribose) polymerase turnover alterations do not contribute to PARP overexpression in Ewing's sarcoma cells". Oncol. Rep. 9 (3): 529–32. doi:10.3892/or.9.3.529. PMID 11956622.
  37. ^ Singh P, Yang M, Dai H, Yu D, Huang Q, Tan W, Kernstine KH, Lin D, Shen B (2008). "Overexpression and hypomethylation of flap endonuclease 1 gene in breast and other cancers". Mol. Cancer Res. 6 (11): 1710–7. doi:10.1158/1541-7786.MCR-08-0269. PMC 2948671. PMID 19010819.
  38. ^ Lam JS, Seligson DB, Yu H, Li A, Eeva M, Pantuck AJ, Zeng G, Horvath S, Belldegrun AS (2006). "Flap endonuclease 1 is overexpressed in prostate cancer and is associated with a high Gleason score". BJU Int. 98 (2): 445–51. doi:10.1111/j.1464-410X.2006.06224.x. PMID 16879693. S2CID 22165252.
  39. ^ Kim JM, Sohn HY, Yoon SY, Oh JH, Yang JO, Kim JH, et al. (2005). "Identification of gastric cancer-related genes using a cDNA microarray containing novel expressed sequence tags expressed in gastric cancer cells". Clin. Cancer Res. 11 (2 Pt 1): 473–82. doi:10.1158/1078-0432.473.11.2. PMID 15701830.
  40. ^ Wang K, Xie C, Chen D (2014). "Flap endonuclease 1 is a promising candidate biomarker in gastric cancer and is involved in cell proliferation and apoptosis". Int. J. Mol. Med. 33 (5): 1268–74. doi:10.3892/ijmm.2014.1682. PMID 24590400.
  41. ^ Krause A, Combaret V, Iacono I, Lacroix B, Compagnon C, Bergeron C, et al. (2005). "Genome-wide analysis of gene expression in neuroblastomas detected by mass screening" (PDF). Cancer Lett. 225 (1): 111–20. doi:10.1016/j.canlet.2004.10.035. PMID 15922863. S2CID 44644467.
  42. ^ Iacobuzio-Donahue CA, Maitra A, Olsen M, Lowe AW, van Heek NT, Rosty C, et al. (2003). "Exploration of global gene expression patterns in pancreatic adenocarcinoma using cDNA microarrays". Am. J. Pathol. 162 (4): 1151–62. doi:10.1016/S0002-9440(10)63911-9. PMC 1851213. PMID 12651607.
  43. ^ Nikolova T, Christmann M, Kaina B (2009). "FEN1 is overexpressed in testis, lung and brain tumors". Anticancer Res. 29 (7): 2453–9. PMID 19596913.
  44. ^ Kastan MB (2008). "DNA damage responses: mechanisms and roles in human disease: 2007 G.H.A. Clowes Memorial Award Lecture". Mol. Cancer Res. 6 (4): 517–24. doi:10.1158/1541-7786.MCR-08-0020. PMID 18403632.
  45. ^ Bernstein, C; Prasad, AR; Nfonsam, V; Bernstein, H. (2013). "Chapter 16: DNA Damage, DNA Repair and Cancer". In Chen, Clark (ed.). New Research Directions in DNA Repair. p. 413. ISBN 978-953-51-1114-6.
  46. ^ O'Hagan HM, Mohammad HP, Baylin SB (2008). "Double strand breaks can initiate gene silencing and SIRT1-dependent onset of DNA methylation in an exogenous promoter CpG island". PLOS Genetics. 4 (8): e1000155. doi:10.1371/journal.pgen.1000155. PMC 2491723. PMID 18704159.
  47. ^ Cuozzo C, Porcellini A, Angrisano T, et al. (July 2007). "DNA damage, homology-directed repair, and DNA methylation". PLOS Genetics. 3 (7): e110. doi:10.1371/journal.pgen.0030110. PMC 1913100. PMID 17616978.
  48. ^ Horvath S (2013). "DNA methylation age of human tissues and cell types". Genome Biology. 14 (10): R115. doi:10.1186/gb-2013-14-10-r115. PMC 4015143. PMID 24138928.
  49. ^ Ramirez-Ortiz ZG, Specht CA, Wang JP, Lee CK, Bartholomeu DC, Gazzinelli RT, Levitz SM (2008). "Toll-like receptor 9-dependent immune activation by unmethylated CpG motifs in Aspergillus fumigatus DNA". Infect. Immun. 76 (5): 2123–2129. doi:10.1128/IAI.00047-08. PMC 2346696. PMID 18332208.
  50. ^ Ziller MJ, Müller F, Liao J, Zhang Y, Gu H, Bock C, et al. (December 2011). "Genomic distribution and inter-sample variation of non-CpG methylation across human cell types". PLOS Genet. 7 (12): e1002389. doi:10.1371/journal.pgen.1002389. PMC 3234221. PMID 22174693.
  51. ^ a b Fasolino M, Zhou Z (May 2017). "The Crucial Role of DNA Methylation and MeCP2 in Neuronal Function". Genes (Basel). 8 (5): 141. doi:10.3390/genes8050141. PMC 5448015. PMID 28505093.
  52. ^ a b Duke CG, Kennedy AJ, Gavin CF, Day JJ, Sweatt JD (July 2017). "Experience-dependent epigenomic reorganization in the hippocampus". Learn. Mem. 24 (7): 278–288. doi:10.1101/lm.045112.117. PMC 5473107. PMID 28620075.
  53. ^ Halder R, Hennion M, Vidal RO, Shomroni O, Rahman RU, Rajput A, et al. (January 2016). "DNA methylation changes in plasticity genes accompany the formation and maintenance of memory". Nat. Neurosci. 19 (1): 102–10. doi:10.1038/nn.4194. PMC 4700510. PMID 26656643.
  54. ^ a b c Zhou X, Zhuang Z, Wang W, He L, Wu H, Cao Y, Pan F, Zhao J, Hu Z, Sekhar C, Guo Z (September 2016). "OGG1 is essential in oxidative stress induced DNA demethylation". Cell. Signal. 28 (9): 1163–71. doi:10.1016/j.cellsig.2016.05.021. PMID 27251462.
  55. ^ Bayraktar G, Kreutz MR (2018). "The Role of Activity-Dependent DNA Demethylation in the Adult Brain and in Neurological Disorders". Front Mol Neurosci. 11: 169. doi:10.3389/fnmol.2018.00169. PMC 5975432. PMID 29875631.
  56. ^ Massaad CA, Klann E (May 2011). "Reactive oxygen species in the regulation of synaptic plasticity and memory". Antioxid. Redox Signal. 14 (10): 2013–54. doi:10.1089/ars.2010.3208. PMC 3078504. PMID 20649473.
  57. ^ Beckhauser TF, Francis-Oliveira J, De Pasquale R (2016). "Reactive Oxygen Species: Physiological and Physiopathological Effects on Synaptic Plasticity". J Exp Neurosci. 10 (Suppl 1): 23–48. doi:10.4137/JEN.S39887. PMC 5012454. PMID 27625575.
  58. ^ Day JJ, Sweatt JD (November 2010). "DNA methylation and memory formation". Nat. Neurosci. 13 (11): 1319–23. doi:10.1038/nn.2666. PMC 3130618. PMID 20975755.
  59. ^ a b c Zhou, Wanding; Liang, Gangning; Molloy, Peter L.; Jones, Peter A. (11 August 2020). "DNA methylation enables transposable element-driven genome expansion". Proceedings of the National Academy of Sciences of the United States of America. 117 (32): 19359–19366. Bibcode:2020PNAS..11719359Z. doi:10.1073/pnas.1921719117. ISSN 1091-6490. PMC 7431005. PMID 32719115.{{cite journal}}: CS1 maint: date and year (link)
  60. ^ Zhou, Wanding; Liang, Gangning; Molloy, Peter L.; Jones, Peter A. (11 August 2020). "DNA methylation enables transposable element-driven genome expansion". Proceedings of the National Academy of Sciences of the United States of America. 117 (32): 19359–19366. Bibcode:2020PNAS..11719359Z. doi:10.1073/pnas.1921719117. ISSN 1091-6490. PMC 7431005. PMID 32719115.

site, confused, with, oligodeoxynucleotide, sites, regions, where, cytosine, nucleotide, followed, guanine, nucleotide, linear, sequence, bases, along, direction, occur, with, high, frequency, genomic, regions, called, islands, islands, phosphate, sequence, nu. Not to be confused with CpG oligodeoxynucleotide The CpG sites or CG sites are regions of DNA where a cytosine nucleotide is followed by a guanine nucleotide in the linear sequence of bases along its 5 3 direction CpG sites occur with high frequency in genomic regions called CpG islands or CG islands a CpG site i e the 5 C phosphate G 3 sequence of nucleotides is indicated on one DNA strand in yellow On the reverse DNA strand in blue the complementary 5 CpG 3 site is shown A C G base pairing between the two DNA strands is also indicated right Cytosines in CpG dinucleotides can be methylated to form 5 methylcytosines Enzymes that add a methyl group are called DNA methyltransferases In mammals 70 to 80 of CpG cytosines are methylated 1 Methylating the cytosine within a gene can change its expression a mechanism that is part of a larger field of science studying gene regulation that is called epigenetics Methylated cytosines often mutate to thymines In humans about 70 of promoters located near the transcription start site of a gene proximal promoters contain a CpG island 2 3 Contents 1 CpG characteristics 1 1 Definition 1 2 Under representation caused by high mutation rate 1 3 Genomic distribution 2 CpG islands 3 Methylation silencing cancer and aging 3 1 CpG islands in promoters 3 2 Methylation of CpG islands stably silences genes 3 3 Promoter CpG hyper hypo methylation in cancer 3 4 DNA repair genes with hyper hypo methylated promoters in cancers 3 5 Methylation of CpG sites with age 3 6 Unmethylated sites 4 Role of CpG sites in memory 4 1 Demethylation at CpG sites requires ROS activity 5 CpG loss 5 1 Genome size and CpG ratio are negatively correlated 5 1 1 Alu elements as promoters of CpG loss 6 See also 7 ReferencesCpG characteristics EditDefinition Edit CpG is shorthand for 5 C phosphate G 3 that is cytosine and guanine separated by only one phosphate group phosphate links any two nucleosides together in DNA The CpG notation is used to distinguish this single stranded linear sequence from the CG base pairing of cytosine and guanine for double stranded sequences The CpG notation is therefore to be interpreted as the cytosine being 5 prime to the guanine base CpG should not be confused with GpC the latter meaning that a guanine is followed by a cytosine in the 5 3 direction of a single stranded sequence Under representation caused by high mutation rate Edit CpG dinucleotides have long been observed to occur with a much lower frequency in the sequence of vertebrate genomes than would be expected due to random chance For example in the human genome which has a 42 GC content 4 a pair of nucleotides consisting of cytosine followed by guanine would be expected to occur 0 21 0 21 4 41 displaystyle 0 21 times 0 21 4 41 of the time The frequency of CpG dinucleotides in human genomes is less than one fifth of the expected frequency 5 This underrepresentation is a consequence of the high mutation rate of methylated CpG sites the spontaneously occurring deamination of a methylated cytosine results in a thymine and the resulting G T mismatched bases are often improperly resolved to A T whereas the deamination of unmethylated cytosine results in a uracil which as a foreign base is quickly replaced by a cytosine by the base excision repair mechanism The C to T transition rate at methylated CpG sites is 10 fold higher than at unmethylated sites 6 7 8 9 Genomic distribution Edit CpG sites GpC sites Distribution of CpG sites left in red and GpC sites right in green in the human APRT gene CpG are more abundant in the upstream region of the gene where they form a CpG island whereas GpC are more evenly distributed The 5 exons of the APRT gene are indicated blue and the start ATG and stop TGA codons are emphasized bold blue CpG dinucleotides frequently occur in CpG islands see definition of CpG islands below There are 28 890 CpG islands in the human genome 50 267 if one includes CpG islands in repeat sequences 10 This is in agreement with the 28 519 CpG islands found by Venter et al 11 since the Venter et al genome sequence did not include the interiors of highly similar repetitive elements and the extremely dense repeat regions near the centromeres 12 Since CpG islands contain multiple CpG dinucleotide sequences there appear to be more than 20 million CpG dinucleotides in the human genome CpG islands Edit How methylation of CpG sites followed by spontaneous deamination leads to a lack of CpG sites in methylated DNA As a result residual CpG islands are created in areas where methylation is rare and CpG sites stick or where C to T mutation is highly detrimental CpG islands or CG islands are regions with a high frequency of CpG sites Though objective definitions for CpG islands are limited the usual formal definition is a region with at least 200 bp a GC percentage greater than 50 and an observed to expected CpG ratio greater than 60 The observed to expected CpG ratio can be derived where the observed is calculated as number of C p G s displaystyle text number of CpGs and the expected as number of C number of G length of sequence displaystyle text number of C text number of G text length of sequence 13 or number of C number of G 2 2 length of sequence displaystyle text number of C text number of G 2 2 text length of sequence 14 Many genes in mammalian genomes have CpG islands associated with the start of the gene 15 promoter regions Because of this the presence of a CpG island is used to help in the prediction and annotation of genes In mammalian genomes CpG islands are typically 300 3 000 base pairs in length and have been found in or near approximately 40 of promoters of mammalian genes 16 Over 60 of human genes and almost all house keeping genes have their promoters embedded in CpG islands 17 Given the frequency of GC two nucleotide sequences the number of CpG dinucleotides is much lower than would be expected 14 A 2002 study revised the rules of CpG island prediction to exclude other GC rich genomic sequences such as Alu repeats Based on an extensive search on the complete sequences of human chromosomes 21 and 22 DNA regions greater than 500 bp were found more likely to be the true CpG islands associated with the 5 regions of genes if they had a GC content greater than 55 and an observed to expected CpG ratio of 65 18 CpG islands are characterized by CpG dinucleotide content of at least 60 of that which would be statistically expected 4 6 whereas the rest of the genome has much lower CpG frequency 1 a phenomenon called CG suppression Unlike CpG sites in the coding region of a gene in most instances the CpG sites in the CpG islands of promoters are unmethylated if the genes are expressed This observation led to the speculation that methylation of CpG sites in the promoter of a gene may inhibit gene expression Methylation along with histone modification is central to imprinting 19 Most of the methylation differences between tissues or between normal and cancer samples occur a short distance from the CpG islands at CpG island shores rather than in the islands themselves 20 CpG islands typically occur at or near the transcription start site of genes particularly housekeeping genes in vertebrates 14 A C cytosine base followed immediately by a G guanine base a CpG is rare in vertebrate DNA because the cytosines in such an arrangement tend to be methylated This methylation helps distinguish the newly synthesized DNA strand from the parent strand which aids in the final stages of DNA proofreading after duplication However over time methylated cytosines tend to turn into thymines because of spontaneous deamination There is a special enzyme in humans Thymine DNA glycosylase or TDG that specifically replaces T s from T G mismatches However due to the rarity of CpGs it is theorised to be insufficiently effective in preventing a possibly rapid mutation of the dinucleotides The existence of CpG islands is usually explained by the existence of selective forces for relatively high CpG content or low levels of methylation in that genomic area perhaps having to do with the regulation of gene expression A 2011 study showed that most CpG islands are a result of non selective forces 21 Methylation silencing cancer and aging Edit An image showing a hypothetical evolutionary mechanism behind CpG island formation Main article DNA methylation CpG islands in promoters Edit In humans about 70 of promoters located near the transcription start site of a gene proximal promoters contain a CpG island 2 3 Distal promoter elements also frequently contain CpG islands An example is the DNA repair gene ERCC1 where the CpG island containing element is located about 5 400 nucleotides upstream of the transcription start site of the ERCC1 gene 22 CpG islands also occur frequently in promoters for functional noncoding RNAs such as microRNAs 23 Methylation of CpG islands stably silences genes Edit In humans DNA methylation occurs at the 5 position of the pyrimidine ring of the cytosine residues within CpG sites to form 5 methylcytosines The presence of multiple methylated CpG sites in CpG islands of promoters causes stable silencing of genes 24 Silencing of a gene may be initiated by other mechanisms but this is often followed by methylation of CpG sites in the promoter CpG island to cause the stable silencing of the gene 24 Promoter CpG hyper hypo methylation in cancer Edit In cancers loss of expression of genes occurs about 10 times more frequently by hypermethylation of promoter CpG islands than by mutations For example in a colorectal cancer there are usually about 3 to 6 driver mutations and 33 to 66 hitchhiker or passenger mutations 25 In contrast in one study of colon tumors compared to adjacent normal appearing colonic mucosa 1 734 CpG islands were heavily methylated in tumors whereas these CpG islands were not methylated in the adjacent mucosa 26 Half of the CpG islands were in promoters of annotated protein coding genes 26 suggesting that about 867 genes in a colon tumor have lost expression due to CpG island methylation A separate study found an average of 1 549 differentially methylated regions hypermethylated or hypomethylated in the genomes of six colon cancers compared to adjacent mucosa of which 629 were in known promoter regions of genes 27 A third study found more than 2 000 genes differentially methylated between colon cancers and adjacent mucosa Using gene set enrichment analysis 569 out of 938 gene sets were hypermethylated and 369 were hypomethylated in cancers 28 Hypomethylation of CpG islands in promoters results in overexpression of the genes or gene sets affected One 2012 study 29 listed 147 specific genes with colon cancer associated hypermethylated promoters along with the frequency with which these hypermethylations were found in colon cancers At least 10 of those genes had hypermethylated promoters in nearly 100 of colon cancers They also indicated 11 microRNAs whose promoters were hypermethylated in colon cancers at frequencies between 50 and 100 of cancers MicroRNAs miRNAs are small endogenous RNAs that pair with sequences in messenger RNAs to direct post transcriptional repression On average each microRNA represses several hundred target genes 30 Thus microRNAs with hypermethylated promoters may be allowing over expression of hundreds to thousands of genes in a cancer The information above shows that in cancers promoter CpG hyper hypo methylation of genes and of microRNAs causes loss of expression or sometimes increased expression of far more genes than does mutation DNA repair genes with hyper hypo methylated promoters in cancers Edit DNA repair genes are frequently repressed in cancers due to hypermethylation of CpG islands within their promoters In head and neck squamous cell carcinomas at least 15 DNA repair genes have frequently hypermethylated promoters these genes are XRCC1 MLH3 PMS1 RAD51B XRCC3 RAD54B BRCA1 SHFM1 GEN1 FANCE FAAP20 SPRTN SETMAR HUS1 and PER1 31 About seventeen types of cancer are frequently deficient in one or more DNA repair genes due to hypermethylation of their promoters 32 As an example promoter hypermethylation of the DNA repair gene MGMT occurs in 93 of bladder cancers 88 of stomach cancers 74 of thyroid cancers 40 90 of colorectal cancers and 50 of brain cancers Promoter hypermethylation of LIG4 occurs in 82 of colorectal cancers Promoter hypermethylation of NEIL1 occurs in 62 of head and neck cancers and in 42 of non small cell lung cancers Promoter hypermethylation of ATM occurs in 47 of non small cell lung cancers Promoter hypermethylation of MLH1 occurs in 48 of non small cell lung cancer squamous cell carcinomas Promoter hypermethylation of FANCB occurs in 46 of head and neck cancers On the other hand the promoters of two genes PARP1 and FEN1 were hypomethylated and these genes were over expressed in numerous cancers PARP1 and FEN1 are essential genes in the error prone and mutagenic DNA repair pathway microhomology mediated end joining If this pathway is over expressed the excess mutations it causes can lead to cancer PARP1 is over expressed in tyrosine kinase activated leukemias 33 in neuroblastoma 34 in testicular and other germ cell tumors 35 and in Ewing s sarcoma 36 FEN1 is over expressed in the majority of cancers of the breast 37 prostate 38 stomach 39 40 neuroblastomas 41 pancreatic 42 and lung 43 DNA damage appears to be the primary underlying cause of cancer 44 45 If accurate DNA repair is deficient DNA damages tend to accumulate Such excess DNA damage can increase mutational errors during DNA replication due to error prone translesion synthesis Excess DNA damage can also increase epigenetic alterations due to errors during DNA repair 46 47 Such mutations and epigenetic alterations can give rise to cancer see malignant neoplasms Thus CpG island hyper hypo methylation in the promoters of DNA repair genes are likely central to progression to cancer Methylation of CpG sites with age Edit Since age has a strong effect on DNA methylation levels on tens of thousands of CpG sites one can define a highly accurate biological clock referred to as epigenetic clock or DNA methylation age in humans and chimpanzees 48 Unmethylated sites Edit Unmethylated CpG dinucleotide sites can be detected by Toll like receptor 9 TLR 9 49 on plasmacytoid dendritic cells monocytes natural killer NK cells and B cells in humans This is used to detect intracellular viral infection Role of CpG sites in memory EditIn mammals DNA methyltransferases which add methyl groups to DNA bases exhibit a sequence preference for cytosines within CpG sites 50 In the mouse brain 4 2 of all cytosines are methylated primarily in the context of CpG sites forming 5mCpG 51 Most hypermethylated 5mCpG sites increase the repression of associated genes 51 As reviewed by Duke et al neuron DNA methylation repressing expression of particular genes is altered by neuronal activity Neuron DNA methylation is required for synaptic plasticity is modified by experiences and active DNA methylation and demethylation is required for memory formation and maintenance 52 In 2016 Halder et al 53 using mice and in 2017 Duke et al 52 using rats subjected the rodents to contextual fear conditioning causing an especially strong long term memory to form At 24 hours after the conditioning in the hippocampus brain region of rats the expression of 1 048 genes was down regulated usually associated with 5mCpG in gene promoters and the expression of 564 genes was up regulated often associated with hypomethylation of CpG sites in gene promoters At 24 hours after training 9 2 of the genes in the rat genome of hippocampus neurons were differentially methylated However while the hippocampus is essential for learning new information it does not store information itself In the mouse experiments of Halder 1 206 differentially methylated genes were seen in the hippocampus one hour after contextual fear conditioning but these altered methylations were reversed and not seen after four weeks In contrast with the absence of long term CpG methylation changes in the hippocampus substantial differential CpG methylation could be detected in cortical neurons during memory maintenance There were 1 223 differentially methylated genes in the anterior cingulate cortex of mice four weeks after contextual fear conditioning Demethylation at CpG sites requires ROS activity Edit Initiation of DNA demethylation at a CpG site In adult somatic cells DNA methylation typically occurs in the context of CpG dinucleotides CpG sites forming 5 methylcytosine pG or 5mCpG Reactive oxygen species ROS may attack guanine at the dinucleotide site forming 8 hydroxy 2 deoxyguanosine 8 OHdG and resulting in a 5mCp 8 OHdG dinucleotide site The base excision repair enzyme OGG1 targets 8 OHdG and binds to the lesion without immediate excision OGG1 present at a 5mCp 8 OHdG site recruits TET1 and TET1 oxidizes the 5mC adjacent to the 8 OHdG This initiates demethylation of 5mC 54 Demethylation of 5 Methylcytosine 5mC in neuron DNA As reviewed in 2018 55 in brain neurons 5mC is oxidized by the ten eleven translocation TET family of dioxygenases TET1 TET2 TET3 to generate 5 hydroxymethylcytosine 5hmC In successive steps TET enzymes further hydroxylate 5hmC to generate 5 formylcytosine 5fC and 5 carboxylcytosine 5caC Thymine DNA glycosylase TDG recognizes the intermediate bases 5fC and 5caC and excises the glycosidic bond resulting in an apyrimidinic site AP site In an alternative oxidative deamination pathway 5hmC can be oxidatively deaminated by activity induced cytidine deaminase apolipoprotein B mRNA editing complex AID APOBEC deaminases to form 5 hydroxymethyluracil 5hmU or 5mC can be converted to thymine Thy 5hmU can be cleaved by TDG single strand selective monofunctional uracil DNA glycosylase 1 SMUG1 Nei Like DNA Glycosylase 1 NEIL1 or methyl CpG binding protein 4 MBD4 AP sites and T G mismatches are then repaired by base excision repair BER enzymes to yield cytosine Cyt Two reviews 56 57 summarize the large body of evidence for the critical and essential role of ROS in memory formation The DNA demethylation of thousands of CpG sites during memory formation depends on initiation by ROS In 2016 Zhou et al 54 showed that ROS have a central role in DNA demethylation TET1 is a key enzyme involved in demethylating 5mCpG However TET1 is only able to act on 5mCpG if an ROS has first acted on the guanine to form 8 hydroxy 2 deoxyguanosine 8 OHdG resulting in a 5mCp 8 OHdG dinucleotide see first figure in this section 54 After formation of 5mCp 8 OHdG the base excision repair enzyme OGG1 binds to the 8 OHdG lesion without immediate excision Adherence of OGG1 to the 5mCp 8 OHdG site recruits TET1 allowing TET1 to oxidize the 5mC adjacent to 8 OHdG as shown in the first figure in this section This initiates the demethylation pathway shown in the second figure in this section Altered protein expression in neurons controlled by ROS dependent demethylation of CpG sites in gene promoters within neuron DNA is central to memory formation 58 CpG loss EditCpG depletion has been observed in the process of DNA methylation of Transposable Elements TEs where TEs are not only responsible in the genome expansion but also CpG loss in a host DNA TEs can be known as methylation centers whereby the methylation process the TEs spreads into the flanking DNA once in the host DNA This spreading might subsequently result in CpG loss over evolutionary time Older evolutionary times show a higher CpG loss in the flanking DNA compared to the younger evolutionary times Therefore the DNA methylation can lead eventually to the noticeably loss of CpG sites in neighboring DNA 59 Genome size and CpG ratio are negatively correlated Edit CpG methylation contributes to the genome expansion and consequently to CpG depletion This picture shows a genome with no TEs and unmethylated CpG sites and the insertion and transposition of a TE lead to methylation and silencing of the TE Through the process of CpG methylation a decrease in CpG is found 60 Previous studies have confirmed the variety of genomes sizes amount species where invertebrates and vertebrates have small and big genomes compared to humans The genome size is strongly connected to the number of transposable elements However there is a correlation between the number of TEs methylation versus the CpG amount This negative correlation consequently causes depletion of CpG due to intergenic DNA methylation which is mostly attributed to the methylation of TEs Overall this contributes to a noticeable amount of CpG loss in different genomes species 59 Alu elements as promoters of CpG loss Edit Alu elements are known as the most abundant type of transposable elements Some studies have used Alu elements as a way to study the idea of which factor is responsible for genome expansion Alu elements are CpG rich in a longer amount of sequence unlike LINEs and ERVs Alus can work as a methylation center and the insertion into a host DNA can produce DNA methylation and provoke a spreading into the Flanking DNA area This spreading is why there are a considerable amount CpG loss and a considerable increase in genome expansion 59 However this is a result that is analyzed over time because older Alus elements show more CpG loss in sites of neighboring DNA compared to younger ones See also EditTLR9 detector of unmethylated CpG sites DNA methylation ageReferences Edit Jabbari K Bernardi G May 2004 Cytosine methylation and CpG TpG CpA and TpA frequencies Gene 333 143 9 doi 10 1016 j gene 2004 02 043 PMID 15177689 a b Saxonov S Berg P Brutlag DL 2006 A genome wide analysis of CpG dinucleotides in the human genome distinguishes two distinct classes of promoters Proc Natl Acad Sci U S A 103 5 1412 7 Bibcode 2006PNAS 103 1412S doi 10 1073 pnas 0510310103 PMC 1345710 PMID 16432200 a b Deaton AM Bird A 2011 CpG islands and the regulation of transcription Genes Dev 25 10 1010 22 doi 10 1101 gad 2037511 PMC 3093116 PMID 21576262 Lander Eric S Linton Lauren M Birren Bruce Nusbaum Chad Zody Michael C Baldwin Jennifer Devon Keri Dewar Ken Doyle Michael 15 February 2001 Initial sequencing and analysis of the human genome Nature 409 6822 860 921 Bibcode 2001Natur 409 860L doi 10 1038 35057062 ISSN 1476 4687 PMID 11237011 International Human Genome Sequencing Consortium 2001 02 15 Initial sequencing and analysis of the human genome Nature 409 6822 860 921 Bibcode 2001Natur 409 860L doi 10 1038 35057062 ISSN 0028 0836 PMID 11237011 Hwang DG Green P 2004 Bayesian Markov chain Monte Carlo sequence analysis reveals varying neutral substitution patterns in mammalian evolution Proc Natl Acad Sci U S A 101 39 13994 4001 Bibcode 2004PNAS 10113994H doi 10 1073 pnas 0404142101 PMC 521089 PMID 15292512 Walsh CP Xu GL 2006 Cytosine methylation and DNA repair Curr Top Microbiol Immunol Current Topics in Microbiology and Immunology 301 283 315 doi 10 1007 3 540 31390 7 11 ISBN 3 540 29114 8 PMID 16570853 Arnheim N Calabrese P 2009 Understanding what determines the frequency and pattern of human germline mutations Nat Rev Genet 10 7 478 488 doi 10 1038 nrg2529 PMC 2744436 PMID 19488047 Segurel L Wyman MJ Przeworski M 2014 Understanding what determines the frequency and pattern of human germline mutations Annu Rev Genom Hum Genet 15 47 70 doi 10 1146 annurev genom 031714 125740 PMID 25000986 Lander ES Linton LM Birren B Nusbaum C Zody MC Baldwin J et al February 2001 Initial sequencing and analysis of the human genome Nature 409 6822 860 921 Bibcode 2001Natur 409 860L doi 10 1038 35057062 PMID 11237011 Venter JC Adams MD Myers EW Li PW Mural RJ Sutton GG et al February 2001 The sequence of the human genome Science 291 5507 1304 51 Bibcode 2001Sci 291 1304V doi 10 1126 science 1058040 PMID 11181995 Myers EW Sutton GG Smith HO Adams MD Venter JC April 2002 On the sequencing and assembly of the human genome Proc Natl Acad Sci U S A 99 7 4145 6 Bibcode 2002PNAS 99 4145M doi 10 1073 pnas 092136699 PMC 123615 PMID 11904395 Gardiner Garden M Frommer M 1987 CpG islands in vertebrate genomes Journal of Molecular Biology 196 2 261 282 doi 10 1016 0022 2836 87 90689 9 PMID 3656447 a b c Saxonov S Berg P Brutlag DL 2006 A genome wide analysis of CpG dinucleotides in the human genome distinguishes two distinct classes of promoters Proc Natl Acad Sci USA 103 5 1412 1417 Bibcode 2006PNAS 103 1412S doi 10 1073 pnas 0510310103 PMC 1345710 PMID 16432200 Hartl DL Jones EW 2005 Genetics Analysis of Genes and Genomes 6th ed Missisauga Jones amp Bartlett Canada p 477 ISBN 978 0 7637 1511 3 Fatemi M Pao MM Jeong S Gal Yam EN Egger G Weisenberger DJ et al 2005 Footprinting of mammalian promoters use of a CpG DNA methyltransferase revealing nucleosome positions at a single molecule level Nucleic Acids Res 33 20 e176 doi 10 1093 nar gni180 PMC 1292996 PMID 16314307 Alberts Bruce 18 November 2014 Molecular biology of the cell Sixth ed New York NY p 406 ISBN 978 0 8153 4432 2 OCLC 887605755 a href Template Cite book html title Template Cite book cite book a CS1 maint location missing publisher link Takai D Jones PA 2002 Comprehensive analysis of CpG islands in human chromosomes 21 and 22 Proc Natl Acad Sci USA 99 6 3740 5 Bibcode 2002PNAS 99 3740T doi 10 1073 pnas 052410099 PMC 122594 PMID 11891299 Feil R Berger F 2007 Convergent evolution of genomic imprinting in plants and mammals Trends Genet 23 4 192 199 doi 10 1016 j tig 2007 02 004 PMID 17316885 Irizarry RA Ladd Acosta C Wen B Wu Z Montano C Onyango P et al 2009 The human colon cancer methylome shows similar hypo and hypermethylation at conserved tissue specific CpG island shores Nature Genetics 41 2 178 186 doi 10 1038 ng 298 PMC 2729128 PMID 19151715 Cohen N Kenigsberg E Tanay A 2011 Primate CpG Islands Are Maintained by Heterogeneous Evolutionary Regimes Involving Minimal Selection Cell 145 5 773 786 doi 10 1016 j cell 2011 04 024 PMID 21620139 S2CID 14856605 Chen HY Shao CJ Chen FR Kwan AL Chen ZP 2010 Role of ERCC1 promoter hypermethylation in drug resistance to cisplatin in human gliomas Int J Cancer 126 8 1944 54 doi 10 1002 ijc 24772 PMID 19626585 Kaur S Lotsari Salomaa JE Seppanen Kaijansinkko R Peltomaki P 2016 MicroRNA Methylation in Colorectal Cancer Adv Exp Med Biol Advances in Experimental Medicine and Biology 937 109 22 doi 10 1007 978 3 319 42059 2 6 ISBN 978 3 319 42057 8 PMID 27573897 a b Bird A 2002 DNA methylation patterns and epigenetic memory Genes Dev 16 1 6 21 doi 10 1101 gad 947102 PMID 11782440 Vogelstein B Papadopoulos N Velculescu VE Zhou S Diaz LA Kinzler KW 2013 Cancer genome landscapes Science 339 6127 1546 58 Bibcode 2013Sci 339 1546V doi 10 1126 science 1235122 PMC 3749880 PMID 23539594 a b Illingworth RS Gruenewald Schneider U Webb S Kerr AR James KD Turner DJ Smith C Harrison DJ Andrews R Bird AP 2010 Orphan CpG islands identify numerous conserved promoters in the mammalian genome PLOS Genet 6 9 e1001134 doi 10 1371 journal pgen 1001134 PMC 2944787 PMID 20885785 Wei J Li G Dang S Zhou Y Zeng K Liu M 2016 Discovery and Validation of Hypermethylated Markers for Colorectal Cancer Dis Markers 2016 1 7 doi 10 1155 2016 2192853 PMC 4963574 PMID 27493446 Beggs AD Jones A El Bahrawy M El Bahwary M Abulafi M Hodgson SV et al 2013 Whole genome methylation analysis of benign and malignant colorectal tumours J Pathol 229 5 697 704 doi 10 1002 path 4132 PMC 3619233 PMID 23096130 Schnekenburger M Diederich M 2012 Epigenetics Offer New Horizons for Colorectal Cancer Prevention Curr Colorectal Cancer Rep 8 1 66 81 doi 10 1007 s11888 011 0116 z PMC 3277709 PMID 22389639 Friedman RC Farh KK Burge CB Bartel DP 2009 Most mammalian mRNAs are conserved targets of microRNAs Genome Res 19 1 92 105 doi 10 1101 gr 082701 108 PMC 2612969 PMID 18955434 Rieke DT Ochsenreither S Klinghammer K Seiwert TY Klauschen F Tinhofer I et al 2016 Methylation of RAD51B XRCC3 and other homologous recombination genes is associated with expression of immune checkpoints and an inflammatory signature in squamous cell carcinoma of the head and neck lung and cervix Oncotarget 7 46 75379 75393 doi 10 18632 oncotarget 12211 PMC 5342748 PMID 27683114 Jin B Robertson KD 2013 DNA methyltransferases DNA damage repair and cancer Adv Exp Med Biol Advances in Experimental Medicine and Biology 754 3 29 doi 10 1007 978 1 4419 9967 2 1 ISBN 978 1 4419 9966 5 PMC 3707278 PMID 22956494 Muvarak N Kelley S Robert C Baer MR Perrotti D Gambacorti Passerini C et al 2015 c MYC Generates Repair Errors via Increased Transcription of Alternative NHEJ Factors LIG3 and PARP1 in Tyrosine Kinase Activated Leukemias Mol Cancer Res 13 4 699 712 doi 10 1158 1541 7786 MCR 14 0422 PMC 4398615 PMID 25828893 Newman EA Lu F Bashllari D Wang L Opipari AW Castle VP 2015 Alternative NHEJ Pathway Components Are Therapeutic Targets in High Risk Neuroblastoma Mol Cancer Res 13 3 470 82 doi 10 1158 1541 7786 MCR 14 0337 PMID 25563294 Mego M Cierna Z Svetlovska D Macak D Machalekova K Miskovska V et al 2013 PARP expression in germ cell tumours J Clin Pathol 66 7 607 12 doi 10 1136 jclinpath 2012 201088 PMID 23486608 S2CID 535704 Newman RE Soldatenkov VA Dritschilo A Notario V 2002 Poly ADP ribose polymerase turnover alterations do not contribute to PARP overexpression in Ewing s sarcoma cells Oncol Rep 9 3 529 32 doi 10 3892 or 9 3 529 PMID 11956622 Singh P Yang M Dai H Yu D Huang Q Tan W Kernstine KH Lin D Shen B 2008 Overexpression and hypomethylation of flap endonuclease 1 gene in breast and other cancers Mol Cancer Res 6 11 1710 7 doi 10 1158 1541 7786 MCR 08 0269 PMC 2948671 PMID 19010819 Lam JS Seligson DB Yu H Li A Eeva M Pantuck AJ Zeng G Horvath S Belldegrun AS 2006 Flap endonuclease 1 is overexpressed in prostate cancer and is associated with a high Gleason score BJU Int 98 2 445 51 doi 10 1111 j 1464 410X 2006 06224 x PMID 16879693 S2CID 22165252 Kim JM Sohn HY Yoon SY Oh JH Yang JO Kim JH et al 2005 Identification of gastric cancer related genes using a cDNA microarray containing novel expressed sequence tags expressed in gastric cancer cells Clin Cancer Res 11 2 Pt 1 473 82 doi 10 1158 1078 0432 473 11 2 PMID 15701830 Wang K Xie C Chen D 2014 Flap endonuclease 1 is a promising candidate biomarker in gastric cancer and is involved in cell proliferation and apoptosis Int J Mol Med 33 5 1268 74 doi 10 3892 ijmm 2014 1682 PMID 24590400 Krause A Combaret V Iacono I Lacroix B Compagnon C Bergeron C et al 2005 Genome wide analysis of gene expression in neuroblastomas detected by mass screening PDF Cancer Lett 225 1 111 20 doi 10 1016 j canlet 2004 10 035 PMID 15922863 S2CID 44644467 Iacobuzio Donahue CA Maitra A Olsen M Lowe AW van Heek NT Rosty C et al 2003 Exploration of global gene expression patterns in pancreatic adenocarcinoma using cDNA microarrays Am J Pathol 162 4 1151 62 doi 10 1016 S0002 9440 10 63911 9 PMC 1851213 PMID 12651607 Nikolova T Christmann M Kaina B 2009 FEN1 is overexpressed in testis lung and brain tumors Anticancer Res 29 7 2453 9 PMID 19596913 Kastan MB 2008 DNA damage responses mechanisms and roles in human disease 2007 G H A Clowes Memorial Award Lecture Mol Cancer Res 6 4 517 24 doi 10 1158 1541 7786 MCR 08 0020 PMID 18403632 Bernstein C Prasad AR Nfonsam V Bernstein H 2013 Chapter 16 DNA Damage DNA Repair and Cancer In Chen Clark ed New Research Directions in DNA Repair p 413 ISBN 978 953 51 1114 6 O Hagan HM Mohammad HP Baylin SB 2008 Double strand breaks can initiate gene silencing and SIRT1 dependent onset of DNA methylation in an exogenous promoter CpG island PLOS Genetics 4 8 e1000155 doi 10 1371 journal pgen 1000155 PMC 2491723 PMID 18704159 Cuozzo C Porcellini A Angrisano T et al July 2007 DNA damage homology directed repair and DNA methylation PLOS Genetics 3 7 e110 doi 10 1371 journal pgen 0030110 PMC 1913100 PMID 17616978 Horvath S 2013 DNA methylation age of human tissues and cell types Genome Biology 14 10 R115 doi 10 1186 gb 2013 14 10 r115 PMC 4015143 PMID 24138928 Ramirez Ortiz ZG Specht CA Wang JP Lee CK Bartholomeu DC Gazzinelli RT Levitz SM 2008 Toll like receptor 9 dependent immune activation by unmethylated CpG motifs in Aspergillus fumigatus DNA Infect Immun 76 5 2123 2129 doi 10 1128 IAI 00047 08 PMC 2346696 PMID 18332208 Ziller MJ Muller F Liao J Zhang Y Gu H Bock C et al December 2011 Genomic distribution and inter sample variation of non CpG methylation across human cell types PLOS Genet 7 12 e1002389 doi 10 1371 journal pgen 1002389 PMC 3234221 PMID 22174693 a b Fasolino M Zhou Z May 2017 The Crucial Role of DNA Methylation and MeCP2 in Neuronal Function Genes Basel 8 5 141 doi 10 3390 genes8050141 PMC 5448015 PMID 28505093 a b Duke CG Kennedy AJ Gavin CF Day JJ Sweatt JD July 2017 Experience dependent epigenomic reorganization in the hippocampus Learn Mem 24 7 278 288 doi 10 1101 lm 045112 117 PMC 5473107 PMID 28620075 Halder R Hennion M Vidal RO Shomroni O Rahman RU Rajput A et al January 2016 DNA methylation changes in plasticity genes accompany the formation and maintenance of memory Nat Neurosci 19 1 102 10 doi 10 1038 nn 4194 PMC 4700510 PMID 26656643 a b c Zhou X Zhuang Z Wang W He L Wu H Cao Y Pan F Zhao J Hu Z Sekhar C Guo Z September 2016 OGG1 is essential in oxidative stress induced DNA demethylation Cell Signal 28 9 1163 71 doi 10 1016 j cellsig 2016 05 021 PMID 27251462 Bayraktar G Kreutz MR 2018 The Role of Activity Dependent DNA Demethylation in the Adult Brain and in Neurological Disorders Front Mol Neurosci 11 169 doi 10 3389 fnmol 2018 00169 PMC 5975432 PMID 29875631 Massaad CA Klann E May 2011 Reactive oxygen species in the regulation of synaptic plasticity and memory Antioxid Redox Signal 14 10 2013 54 doi 10 1089 ars 2010 3208 PMC 3078504 PMID 20649473 Beckhauser TF Francis Oliveira J De Pasquale R 2016 Reactive Oxygen Species Physiological and Physiopathological Effects on Synaptic Plasticity J Exp Neurosci 10 Suppl 1 23 48 doi 10 4137 JEN S39887 PMC 5012454 PMID 27625575 Day JJ Sweatt JD November 2010 DNA methylation and memory formation Nat Neurosci 13 11 1319 23 doi 10 1038 nn 2666 PMC 3130618 PMID 20975755 a b c Zhou Wanding Liang Gangning Molloy Peter L Jones Peter A 11 August 2020 DNA methylation enables transposable element driven genome expansion Proceedings of the National Academy of Sciences of the United States of America 117 32 19359 19366 Bibcode 2020PNAS 11719359Z doi 10 1073 pnas 1921719117 ISSN 1091 6490 PMC 7431005 PMID 32719115 a href Template Cite journal html title Template Cite journal cite journal a CS1 maint date and year link Zhou Wanding Liang Gangning Molloy Peter L Jones Peter A 11 August 2020 DNA methylation enables transposable element driven genome expansion Proceedings of the National Academy of Sciences of the United States of America 117 32 19359 19366 Bibcode 2020PNAS 11719359Z doi 10 1073 pnas 1921719117 ISSN 1091 6490 PMC 7431005 PMID 32719115 Portal Biology Retrieved from https en wikipedia org w index php title CpG site amp oldid 1163307852, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.