Search This Blog

Friday, May 10, 2019

Cancer epigenetics

From Wikipedia, the free encyclopedia

Cancer epigenetics is the study of epigenetic modifications to the DNA of cancer cells that do not involve a change in the nucleotide sequence. Epigenetic alterations may be just as important, or even more important, than genetic mutations in a cell's transformation to cancer. In cancers, loss of expression of genes occurs about 10 times more frequently by transcription silencing (caused by epigenetic promoter hypermethylation of CpG islands) than by mutations. As Vogelstein et al. point out, in a colorectal cancer there are usually about 3 to 6 driver mutations and 33 to 66 hitchhiker or passenger mutations. However, in colon tumors compared to adjacent normal-appearing colonic mucosa, there are about 600 to 800 heavily methylated CpG islands in promoters of genes in the tumors while these CpG islands are not methylated in the adjacent mucosa. Manipulation of epigenetic alterations holds great promise for cancer prevention, detection, and therapy. In different types of cancer, a variety of epigenetic mechanisms can be perturbed, such as silencing of tumor suppressor genes and activation of oncogenes by altered CpG island methylation patterns, histone modifications, and dysregulation of DNA binding proteins. Several medications which have epigenetic impact are now used in several of these diseases.
 
Epigenetics patterns in a normal and cancer cells
 
Epigenetic alterations in tumour progression

Mechanisms

DNA methylation

A DNA molecule fragment that is methylated at two cytosines
 
In somatic cells, patterns of DNA methylation are in general transmitted to daughter cells with high fidelity. However, epigenetic DNA methylation differs between normal cells and tumor cells in humans. The "normal" CpG methylation profile is often inverted in cells that become tumorigenic. In normal cells, CpG islands preceding gene promoters are generally unmethylated, and tend to be transcriptionally active, while other individual CpG dinucleotides throughout the genome tend to be methylated. However, in cancer cells, CpG islands preceding tumor suppressor gene promoters are often hypermethylated, while CpG methylation of oncogene promoter regions and parasitic repeat sequences is often decreased.

Hypermethylation of tumor suppressor gene promoter regions can result in silencing of those genes. This type of epigenetic mutation allows cells to grow and reproduce uncontrollably, leading to tumorigenesis. Genes commonly found to be transcriptionally silenced due to promoter hypermethylation include: Cyclin-dependent kinase inhibitor p16, a cell-cycle inhibitor; MGMT, a DNA repair gene; APC, a cell cycle regulator; MLH1, a DNA-repair gene; and BRCA1, another DNA-repair gene. Indeed, cancer cells can become addicted to the transcriptional silencing, due to promoter hypermethylation, of some key tumor suppressor genes, a process known as epigenetic addiction.

Hypomethylation of CpG dinucleotides in other parts of the genome leads to chromosome instability due to mechanisms such as loss of imprinting and reactivation of transposable elements. Loss of imprinting of insulin-like growth factor gene (IGF2) increases risk of colorectal cancer and is associated with Beckwith-Wiedemann syndrome which significantly increases the risk of cancer for newborns. In healthy cells, CpG dinucleotides of lower densities are found within coding and non-coding intergenic regions. Expression of some repetitive sequences and meiotic recombination at centromeres are repressed through methylation.

The entire genome of a cancerous cell contains significantly less methylcytosine than the genome of a healthy cell. In fact, cancer cell genomes have 20-50% less methylation at individual CpG dinucleotides across the genome. CpG islands found in promoter regions are usually protected from DNA methylation. In cancer cells CpG islands are hypomethylated  The regions flanking CpG islands called CpG island shores are where most DNA methylation occurs in the CpG dinucleotide context. Cancer cells are deferentially methylated at CpG island shores. In cancer cells, hypermethylation in the CpG island shores move into CpG islands, or hypomethylation of CpG islands move into CpG island shores eliminating sharp epigenetic boundaries between these genetic elements. In cancer cells "global hypomethylation" due to disruption in DNA methyltransferases (DNMTs) may promote mitotic recombination and chromosome rearrangement, ultimately resulting in aneuploidy when the chromosomes fail to separate properly during mitosis.

CpG island methylation is important in regulation of gene expression, yet cytosine methylation can lead directly to destabilizing genetic mutations and a precancerous cellular state. Methylated cytosines make hydrolysis of the amine group and spontaneous conversion to thymine more favorable. They can cause aberrant recruitment of chromatin proteins. Cytosine methylations change the amount of UV light absorption of the nucleotide base, creating pyrimidine dimers. When mutation results in loss of heterozygosity at tumor suppressor gene sites, these genes may become inactive. Single base pair mutations during replication can also have detrimental effects.

Histone modification

Eukaryotic DNA has a complex structure. It is generally wrapped around special proteins called histones to form a structure called a nucleosome. A nucleosome consists of 2 sets of 4 histones: H2A, H2B, H3, and H4. Additionally, histone H1 contributes to DNA packaging outside of the nucleosome. Certain histone modifying enzymes can add or remove functional groups to the histones, and these modifications influence the level of transcription of the genes wrapped around those histones and the level of DNA replication. Histone modification profiles of healthy and cancerous cells tend to differ. 

In comparison to healthy cells, cancerous cells exhibit decreased monoacetylated and trimethylated forms of histone H4 (decreased H4ac and H4me3). Additionally, mouse models have shown that a decrease in histone H4R3 asymmetric dimethylation (H4R3me2a) of the p19ARF promoter is correlated with more advanced cases of tumorigenesis and metastasis. In mouse models, the loss of histone H4 acetylation and trimethylation increases as tumor growth continues. Loss of histone H4 Lysine 16 acetylation (H4K16ac), which is a mark of aging at the telomeres, specifically loses its acetylation. Some scientists hope this particular loss of histone acetylation might be battled with a histone deacetylase (HDAC) inhibitor specific for SIRT1, an HDAC specific for H4K16.

Other histone marks associated with tumorigenesis include increased deacetylation (decreased acetylation) of histones H3 and H4, decreased trimethylation of histone H3 Lysine 4 (H3K4me3), and increased monomethylation of histone H3 Lysine 9 (H3K9me) and trimethylation of histone H3 Lysine 27 (H3K27me3). These histone modifications can silence tumor suppressor genes despite the drop in methylation of the gene's CpG island (an event that normally activates genes).

Some research has focused on blocking the action of BRD4 on acetylated histones, which has been shown to increase the expression of the Myc protein, implicated in several cancers. The development process of the drug to bind to BRD4 is noteworthy for the collaborative, open approach the team is taking.

The tumor suppressor gene p53 regulates DNA repair and can induce apoptosis in dysregulated cells. E Soto-Reyes and F Recillas-Targa elucidated the importance of the CTCF protein in regulating p53 expression.[26] CTCF, or CCCTC binding factor, is a zinc finger protein that insulates the p53 promoter from accumulating repressive histone marks. In certain types of cancer cells, the CTCF protein does not bind normally, and the p53 promoter accumulates repressive histone marks, causing p53 expression to decrease.

Mutations in the epigenetic machinery itself may occur as well, potentially responsible for the changing epigenetic profiles of cancerous cells. The histone variants of the H2A family are highly conserved in mammals, playing critical roles in regulating many nuclear processes by altering chromatin structure. One of the key H2A variants, H2A.X, marks DNA damage, facilitating the recruitment of DNA repair proteins to restore genomic integrity. Another variant, H2A.Z, plays an important role in both gene activation and repression. A high level of H2A.Z expression is detected in many cancers and is significantly associated with cellular proliferation and genomic instability. Histone variant macroH2A1 is important in the pathogenesis of many types of cancers, for instance in hepatocellular carcinoma. Other mechanisms include a decrease in H4K16ac may be caused by either a decrease in activity of a histone acetyltransferases (HATs) or an increase in deacetylation by SIRT1. Likewise, an inactivating frameshift mutation in HDAC2, a histone deacetylase that acts on many histone-tail lysines, has been associated with cancers showing altered histone acetylation patterns. These findings indicate a promising mechanism for altering epigenetic profiles through enzymatic inhibition or enhancement. 

DNA damage, caused by UV light, ionizing radiation, environmental toxins, and metabolic chemicals, can also lead to genomic instability and cancer. The DNA damage response to double strand DNA breaks (DSB) is mediated in part by histone modifications. At a DSB, MRE11-RAD50-NBS1 (MRN) protein complex recruits ataxia telangiectasia mutated (ATM) kinase which phosphorylates Serine 129 of Histone 2A. MDC1, mediator of DNA damage checkpoint 1, binds to the phosphopeptide, and phosphorylation of H2AX may spread by a positive feedback loop of MRN-ATM recruitment and phosphorylation. TIP60 acetylates the γH2AX, which is then polyubiquitylated. RAP80, a subunit of the DNA repair breast cancer type 1 susceptibility protein complex (BRCA1-A), binds ubiquitin attached to histones. BRCA1-A activity arrests the cell cycle at the G2/M checkpoint, allowing time for DNA repair, or apoptosis may be initiated.

MicroRNA gene silencing

In mammals, microRNAs (miRNAs) regulate about 60% of the transcriptional activity of protein-encoding genes. Some miRNAs also undergo methylation-associated silencing in cancer cells. Let-7 and miR15/16 play important roles in down-regulating RAS and BCL2 oncogenes, and their silencing occurs in cancer cells. Decreased expression of miR-125b1, a miRNA that functions as a tumor suppressor, was observed in prostate, ovarian, breast and glial cell cancers. In vitro experiments have shown that miR-125b1 targets two genes, HER2/neu and ESR1, that are linked to breast cancer. DNA methylation, specifically hypermethylation, is one of the main ways that the miR-125b1 is epigenetically silenced. In patients with breast cancer, hypermethylation of CpG islands located proximal to the transcription start site was observed. Loss of CTCF binding and an increase in repressive histone marks, H3K9me3 and H3K27me3, correlates with DNA methylation and miR-125b1 silencing. Mechanistically, CTCF may function as a boundary element to stop the spread of DNA methylation. Results from experiments conducted by Soto-Reyes et al. indicate a negative effect of methylation on the function and expression of miR-125b1. Therefore, they concluded that DNA methylation has a part in silencing the gene. Furthermore, some miRNA's are epigenetically silenced early on in breast cancer, and therefore these miRNA's could potentially be useful as tumor markers. The epigenetic silencing of miRNA genes by aberrant DNA methylation is a frequent event in cancer cells; almost one third of miRNA promoters active in normal mammary cells were found hypermethylated in breast cancer cells - that is a several fold greater proportion than is usually observed for protein coding genes.

Metabolic recoding of epigenetics in cancer

Dysregulation of metabolism allows tumor cells to generate needed building blocks as well as to modulate epigenetic marks to support cancer initiation and progression. Cancer-induced metabolic changes alter the epigenetic landscape, especially modifications on histones and DNA, thereby promoting malignant transformation, adaptation to inadequate nutrition, and metastasis. The accumulation of certain metabolites in cancer can target epigenetic enzymes to globally alter the epigenetic landscape. Cancer-related metabolic changes lead to locus-specific recoding of epigenetic marks. Cancer epigenetics can be precisely reprogramed by cellular metabolism through 1) dose-responsive modulation of cancer epigenetics by metabolites; 2) sequence-specific recruitment of metabolic enzymes; and 3) targeting of epigenetic enzymes by nutritional signals.

MicroRNA and DNA repair

DNA damage appears to be the primary underlying cause of cancer. If DNA repair is deficient, DNA damage tends 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. Such mutations and epigenetic alterations can give rise to cancer.

Germ line mutations in DNA repair genes cause only 2–5% of colon cancer cases. However, altered expression of microRNAs, causing DNA repair deficiencies, are frequently associated with cancers and may be an important causal factor for these cancers.

Over-expression of certain miRNAs may directly reduce expression of specific DNA repair proteins. Wan et al. referred to 6 DNA repair genes that are directly targeted by the miRNAs indicated in parentheses: ATM (miR-421), RAD52 (miR-210, miR-373), RAD23B (miR-373), MSH2 (miR-21), BRCA1 (miR-182) and P53 (miR-504, miR-125b). More recently, Tessitore et al. listed further DNA repair genes that are directly targeted by additional miRNAs, including ATM (miR-18a, miR-101), DNA-PK (miR-101), ATR (miR-185), Wip1 (miR-16), MLH1, MSH2 and MSH6 (miR-155), ERCC3 and ERCC4 (miR-192) and UNG2 (mir-16, miR-34c and miR-199a). Of these miRNAs, miR-16, miR-18a, miR-21, miR-34c, miR-125b, miR-101, miR-155, miR-182, miR-185 and miR-192 are among those identified by Schnekenburger and Diederich as over-expressed in colon cancer through epigenetic hypomethylation. Over expression of any one of these miRNAs can cause reduced expression of its target DNA repair gene.

Up to 15% of the MLH1-deficiencies in sporadic colon cancers appeared to be due to over-expression of the microRNA miR-155, which represses MLH1 expression. However, the majority of 68 sporadic colon cancers with reduced expression of the DNA mismatch repair protein MLH1 were found to be deficient due to epigenetic methylation of the CpG island of the MLH1 gene.

In 28% of glioblastomas, the MGMT DNA repair protein is deficient but the MGMT promoter is not methylated. In the glioblastomas without methylated MGMT promoters, the level of microRNA miR-181d is inversely correlated with protein expression of MGMT and the direct target of miR-181d is the MGMT mRNA 3’UTR (the three prime untranslated region of MGMT mRNA). Thus, in 28% of glioblastomas, increased expression of miR-181d and reduced expression of DNA repair enzyme MGMT may be a causal factor. In 29–66% of glioblastomas, DNA repair is deficient due to epigenetic methylation of the MGMT gene, which reduces protein expression of MGMT. 

High mobility group A (HMGA) proteins, characterized by an AT-hook, are small, nonhistone, chromatin-associated proteins that can modulate transcription. MicroRNAs control the expression of HMGA proteins, and these proteins (HMGA1 and HMGA2) are architectural chromatin transcription-controlling elements. Palmieri et al showed that, in normal tissues, HGMA1 and HMGA2 genes are targeted (and thus strongly reduced in expression) by miR-15, miR-16, miR-26a, miR-196a2 and Let-7a

HMGA expression is almost undetectable in differentiated adult tissues but is elevated in many cancers. HGMA proteins are polypeptides of ~100 amino acid residues characterized by a modular sequence organization. These proteins have three highly positively charged regions, termed AT hooks, that bind the minor groove of AT-rich DNA stretches in specific regions of DNA. Human neoplasias, including thyroid, prostatic, cervical, colorectal, pancreatic and ovarian carcinoma, show a strong increase of HMGA1a and HMGA1b proteins. Transgenic mice with HMGA1 targeted to lymphoid cells develop aggressive lymphoma, showing that high HMGA1 expression is not only associated with cancers, but that the HMGA1 gene can act as an oncogene to cause cancer. Baldassarre et al., showed that HMGA1 protein binds to the promoter region of DNA repair gene BRCA1 and inhibits BRCA1 promoter activity. They also showed that while only 11% of breast tumors had hypermethylation of the BRCA1 gene, 82% of aggressive breast cancers have low BRCA1 protein expression, and most of these reductions were due to chromatin remodeling by high levels of HMGA1 protein.

HMGA2 protein specifically targets the promoter of ERCC1, thus reducing expression of this DNA repair gene. ERCC1 protein expression was deficient in 100% of 47 evaluated colon cancers (though the extent to which HGMA2 was involved is unknown).

Palmieri et al. showed that each of the miRNAs that target HMGA genes are drastically reduced in almost all human pituitary adenomas studied, when compared with the normal pituitary gland. Consistent with the down-regulation of these HMGA-targeting miRNAs, an increase in the HMGA1 and HMGA2-specific mRNAs was observed. Three of these microRNAs (miR-16, miR-196a and Let-7a) have methylated promoters and therefore low expression in colon cancer. For two of these, miR-15 and miR-16, the coding regions are epigenetically silenced in cancer due to histone deacetylase activity. When these microRNAs are expressed at a low level, then HMGA1 and HMGA2 proteins are expressed at a high level. HMGA1 and HMGA2 target (reduce expression of) BRCA1 and ERCC1 DNA repair genes. Thus DNA repair can be reduced, likely contributing to cancer progression.

DNA repair pathways

A chart of common DNA damaging agents, examples of lesions they cause in DNA, and pathways used to repair these lesions. Also shown are many of the genes in these pathways, an indication of which genes are epigenetically regulated to have reduced (or increased) expression in various cancers. It also shows genes in the error prone microhomology-mediated end joining pathway with increased expression in various cancers.

The chart in this section shows some frequent DNA damaging agents, examples of DNA lesions they cause, and the pathways that deal with these DNA damages. At least 169 enzymes are either directly employed in DNA repair or influence DNA repair processes. Of these, 83 are directly employed in repairing the 5 types of DNA damages illustrated in the chart. 

Some of the more well studied genes central to these repair processes are shown in the chart. The gene designations shown in red, gray or cyan indicate genes frequently epigenetically altered in various types of cancers. Wikipedia articles on each of the genes highlighted by red, gray or cyan describe the epigenetic alteration(s) and the cancer(s) in which these epimutations are found. Three review articles, and two broad experimental survey articles also document most of these epigenetic DNA repair deficiencies in cancers. 

Red-highlighted genes are frequently reduced or silenced by epigenetic mechanisms in various cancers. When these genes have low or absent expression, DNA damages can accumulate. Replication errors past these damages can lead to increased mutations and, ultimately, cancer. Epigenetic repression of DNA repair genes in accurate DNA repair pathways appear to be central to carcinogenesis.

The two gray-highlighted genes RAD51 and BRCA2, are required for homologous recombinational repair. They are sometimes epigenetically over-expressed and sometimes under-expressed in certain cancers. As indicated in the Wikipedia articles on RAD51 and BRCA2, such cancers ordinarily have epigenetic deficiencies in other DNA repair genes. These repair deficiencies would likely cause increased unrepaired DNA damages. The over-expression of RAD51 and BRCA2 seen in these cancers may reflect selective pressures for compensatory RAD51 or BRCA2 over-expression and increased homologous recombinational repair to at least partially deal with such excess DNA damages. In those cases where RAD51 or BRCA2 are under-expressed, this would itself lead to increased unrepaired DNA damages. Replication errors past these damages could cause increased mutations and cancer, so that under-expression of RAD51 or BRCA2 would be carcinogenic in itself. 

Cyan-highlighted genes are in the microhomology-mediated end joining (MMEJ) pathway and are up-regulated in cancer. MMEJ is an additional error-prone inaccurate repair pathway for double-strand breaks. In MMEJ repair of a double-strand break, an homology of 5-25 complementary base pairs between both paired strands is sufficient to align the strands, but mismatched ends (flaps) are usually present. MMEJ removes the extra nucleotides (flaps) where strands are joined, and then ligates the strands to create an intact DNA double helix. MMEJ almost always involves at least a small deletion, so that it is a mutagenic pathway. FEN1, the flap endonuclease in MMEJ, is epigenetically increased by promoter hypomethylation and is over-expressed in the majority of cancers of the breast, prostate, stomach, neuroblastomas, pancreas, and lung. PARP1 is also over-expressed when its promoter region ETS site is epigenetically hypomethylated, and this contributes to progression to endometrial cancer, BRCA-mutated ovarian cancer, and BRCA-mutated serous ovarian cancer. Other genes in the MMEJ pathway are also over-expressed in a number of cancers (see MMEJ for summary), and are also shown in blue.

Frequencies of epimutations in DNA repair genes

Deficiencies in DNA repair proteins that function in accurate DNA repair pathways increase the risk of mutation. Mutation rates are strongly increased in cells with mutations in DNA mismatch repair or in homologous recombinational repair (HRR). Individuals with inherited mutations in any of 34 DNA repair genes are at increased risk of cancer.

In sporadic cancers, a deficiency in DNA repair is occasionally found to be due to a mutation in a DNA repair gene, but much more frequently reduced or absent expression of DNA repair genes is due to epigenetic alterations that reduce or silence gene expression. For example, for 113 colorectal cancers examined in sequence, only four had a missense mutation in the DNA repair gene MGMT, while the majority had reduced MGMT expression due to methylation of the MGMT promoter region (an epigenetic alteration). Similarly, out of 119 cases of mismatch repair-deficient colorectal cancers that lacked DNA repair gene PMS2 expression, PMS2 protein was deficient in 6 due to mutations in the PMS2 gene, while in 103 cases PMS2 expression was deficient because its pairing partner MLH1 was repressed due to promoter methylation (PMS2 protein is unstable in the absence of MLH1). In the other 10 cases, loss of PMS2 expression was likely due to epigenetic overexpression of the microRNA, miR-155, which down-regulates MLH1.

Epigenetic defects in DNA repair genes are frequent in cancers. In the table, multiple cancers were evaluated for reduced or absent expression of the DNA repair gene of interest, and the frequency shown is the frequency with which the cancers had an epigenetic deficiency of gene expression. Such epigenetic deficiencies likely arise early in carcinogenesis, since they are also frequently found (though at somewhat lower frequency) in the field defect surrounding the cancer from which the cancer likely arose (see Table). 

Frequency of epigenetic reduction in DNA repair gene expression in sporadic cancers and in adjacent field defects
Cancer Gene Frequency in Cancer Frequency in Field Defect
Colorectal MGMT 46% 34%
Colorectal MGMT 47% 11%
Colorectal MGMT 70% 60%
Colorectal MSH2 13% 5%
Colorectal ERCC1 100% 40%
Colorectal PMS2 88% 50%
Colorectal XPF 55% 40%
Head and Neck MGMT 54% 38%
Head and Neck MLH1 33% 25%
Head and Neck MLH1 31% 20%
Stomach MGMT 88% 78%
Stomach MLH1 73% 20%
Esophagus MLH1 77%–100% 23%–79%

It appears that cancers may frequently be initiated by an epigenetic reduction in expression of one or more DNA repair enzymes. Reduced DNA repair likely allows accumulation of DNA damages. Error prone translesion synthesis past some of these DNA damages may give rise to a mutation with a selective advantage. A clonal patch with a selective advantage may grow and out-compete neighboring cells, forming a field defect. While there is no obvious selective advantage for a cell to have reduced DNA repair, the epimutation of the DNA repair gene may be carried along as a passenger when the cells with the selectively advantageous mutation are replicated. In the cells carrying both the epimutation of the DNA repair gene and the mutation with the selective advantage, further DNA damages will accumulate, and these could, in turn, give rise to further mutations with still greater selective advantages. Epigenetic defects in DNA repair may thus contribute to the characteristic high frequency of mutations in the genomes of cancers, and cause their carcinogenic progression. 

Cancers have high levels of genome instability, associated with a high frequency of mutations. A high frequency of genomic mutations increases the likelihood of particular mutations occurring that activate oncogenes and inactivate tumor suppressor genes, leading to carcinogenesis. On the basis of whole genome sequencing, cancers are found to have thousands to hundreds of thousands of mutations in their whole genomes. By comparison, the mutation frequency in the whole genome between generations for humans (parent to child) is about 70 new mutations per generation. In the protein coding regions of the genome, there are only about 0.35 mutations between parent/child generations (less than one mutated protein per generation). Whole genome sequencing in blood cells for a pair of identical twin 100-year-old centenarians only found 8 somatic differences, though somatic variation occurring in less than 20% of blood cells would be undetected.

While DNA damages may give rise to mutations through error prone translesion synthesis, DNA damages can also give rise to epigenetic alterations during faulty DNA repair processes. The DNA damages that accumulate due to epigenetic DNA repair defects can be a source of the increased epigenetic alterations found in many genes in cancers. In an early study, looking at a limited set of transcriptional promoters, Fernandez et al. examined the DNA methylation profiles of 855 primary tumors. Comparing each tumor type with its corresponding normal tissue, 729 CpG island sites (55% of the 1322 CpG sites evaluated) showed differential DNA methylation. Of these sites, 496 were hypermethylated (repressed) and 233 were hypomethylated (activated). Thus, there is a high level of epigenetic promoter methylation alterations in tumors. Some of these epigenetic alterations may contribute to cancer progression.

Epigenetic carcinogens

A variety of compounds are considered as epigenetic carcinogens—they result in an increased incidence of tumors, but they do not show mutagen activity (toxic compounds or pathogens that cause tumors incident to increased regeneration should also be excluded). Examples include diethylstilbestrol, arsenite, hexachlorobenzene, and nickel compounds.

Many teratogens exert specific effects on the fetus by epigenetic mechanisms. While epigenetic effects may preserve the effect of a teratogen such as diethylstilbestrol throughout the life of an affected child, the possibility of birth defects resulting from exposure of fathers or in second and succeeding generations of offspring has generally been rejected on theoretical grounds and for lack of evidence. However, a range of male-mediated abnormalities have been demonstrated, and more are likely to exist. FDA label information for Vidaza, a formulation of 5-azacitidine (an unmethylatable analog of cytidine that causes hypomethylation when incorporated into DNA) states that "men should be advised not to father a child" while using the drug, citing evidence in treated male mice of reduced fertility, increased embryo loss, and abnormal embryo development. In rats, endocrine differences were observed in offspring of males exposed to morphine. In mice, second generation effects of diethylstilbesterol have been described occurring by epigenetic mechanisms.

Cancer subtypes

Skin cancer

Melanoma is a deadly skin cancer that originates from melanocytes. Several epigenetic alterations are known to play a role in the transition of melanocytes to melanoma cells. These alterations are the consequence of deregulation of their corresponding enzymes. Several histone methyltransferases and demethylases are among these enzymes.

Prostate cancer

Prostate cancer kills around 35,000 men yearly, and about 220,000 men are diagnosed with prostate cancer per year, in North America alone. Prostate cancer is the second leading cause of cancer-caused fatalities in men, and within a man's lifetime, one in six men will have the disease. Alterations in histone acetylation and DNA methylation occur in various genes influencing prostate cancer. More than 90% of prostate cancers show gene silencing by CpG island hypermethylation of the GSTP1 gene promoter, which protects prostate cells from genomic damage that is caused by different oxidants or carcinogens. Real-time methylation-specific polymerase chain reaction (PCR) suggests that many other genes are also hypermethylated. Gene expression in the prostate may be modulated by nutrition and lifestyle changes.

Cervical cancer

The second most common malignant tumor in women is invasive cervical cancer (ICC) and more than 50% of all invasive cervical cancer (ICC) is caused by oncongenic human papillomavirus 16 (HPV16). Furthermore, cervix intraepithelial neoplasia (CIN) is primarily caused by oncogenic HPV16. As in many cases, the causative factor for cancer does not always take a direct route from infection to the development of cancer. Genomic methylation patterns have been associated with invasive cervical cancer. Within the HPV16L1 region, 14 tested CpG sites have significantly higher methylation in CIN3+ than in HPV16 genomes of women without CIN3. Only 2/16 CpG sites tested in HPV16 upstream regulatory region were found to have association with increased methylation in CIN3+. This suggests that the direct route from infection to cancer is sometimes detoured to a precancerous state in cervix intraepithelial neoplasia. Additionally, increased CpG site methylation was found in low levels in most of the five host nuclear genes studied, including 5/5 TERT, 1/4 DAPK1, 2/5 RARB, MAL, and CADM1. Furthermore, 1/3 of CpG sites in mitochondrial DNA were associated with increased methylation in CIN3+. Thus, a correlation exists between CIN3+ and increased methylation of CpG sites in the HPV16 L1 open reading frame. This could be a potential biomarker for future screens of cancerous and precancerous cervical disease.

Leukemia

Recent studies have shown that the mixed-lineage leukemia (MLL) gene causes leukemia by rearranging and fusing with other genes in different chromosomes, which is a process under epigenetic control.

Sarcoma

There are about 15,000 new cases of sarcoma in the US each year, and about 6,200 people were projected to die of sarcoma in the US in 2014. Sarcomas comprise a large number of rare, histogenetically heterogeneous mesenchymal tumors that, for example, include chondrosarcoma, Ewing's sarcoma, leiomyosarcoma, liposarcoma, osteosarcoma, synovial sarcoma, and (alveolar and embryonal) rhabdomyosarcoma. Several oncogenes and tumor suppressor genes are epigenetically altered in sarcomas. These include APC, CDKN1A, CDKN2A, CDKN2B, Ezrin, FGFR1, GADD45A, MGMT, STK3, STK4, PTEN, RASSF1A, WIF1, as well as several miRNAs. Expression of epigenetic modifiers such as that of the BMI1 component of the PRC1 complex is deregulated in chondrosarcoma, Ewing's sarcoma, and osteosarcoma, and expression of the EZH2 component of the PRC2 complex is altered in Ewing's sarcoma and rhabdomyosarcoma. Similarly, expression of another epigenetic modifier, the LSD1 histone demethylase, is increased in chondrosarcoma, Ewing's sarcoma, osteosarcoma, and rhabdomyosarcoma. Drug targeting and inhibition of EZH2 in Ewing's sarcoma, or of LSD1 in several sarcomas, inhibits tumor cell growth in these sarcomas.

Identification methods

Previously, epigenetic profiles were limited to individual genes under scrutiny by a particular research team. Recently, however, scientists have been moving toward a more genomic approach to determine an entire genomic profile for cancerous versus healthy cells.

Popular approaches for measuring CpG methylation in cells include:
Since bisulfite sequencing is considered the gold standard for measuring CpG methylation, when one of the other methods is used, results are usually confirmed using bisulfite sequencing. Popular approaches for determining histone modification profiles in cancerous versus healthy cells include:

Diagnosis and prognosis

Researchers are hoping to identify specific epigenetic profiles of various types and subtypes of cancer with the goal of using these profiles as tools to diagnose individuals more specifically and accurately. Since epigenetic profiles change, scientists would like to use the different epigenomic profiles to determine the stage of development or level of aggressiveness of a particular cancer in patients. For example, hypermethylation of the genes coding for Death-Associated Protein Kinase (DAPK), p16, and Epithelial Membrane Protein 3 (EMP3) have been linked to more aggressive forms of lung, colorectal, and brain cancers. This type of knowledge can affect the way that doctors will diagnose and choose to treat their patients.

Another factor that will influence the treatment of patients is knowing how well they will respond to certain treatments. Personalized epigenomic profiles of cancerous cells can provide insight into this field. For example, MGMT is an enzyme that reverses the addition of alkyl groups to the nucleotide guanine. Alkylating guanine, however, is the mechanism by which several chemotherapeutic drugs act in order to disrupt DNA and cause cell death. Therefore, if the gene encoding MGMT in cancer cells is hypermethylated and in effect silenced or repressed, the chemotherapeutic drugs that act by methylating guanine will be more effective than in cancer cells that have a functional MGMT enzyme.

Epigenetic biomarkers can also be utilized as tools for molecular prognosis. In primary tumor and mediastinal lymph node biopsy samples, hypermethylation of both CDKN2A and CDH13 serves as the marker for increased risk of faster cancer relapse and higher death rate of patients.

Treatment

Epigenetic control of the proto-onco regions and the tumor suppressor sequences by conformational changes in histones plays a role in the formation and progression of cancer. Pharmaceuticals that reverse epigenetic changes might have a role in a variety of cancers.

Recently, it is evidently known that associations between specific cancer histotypes and epigenetic changes can facilitate the development of novel epi-drugs. Drug development has focused mainly on modifying DNA methyltransferase, histone acetyltransferase (HAT) and histone deacetylase (HDAC).

Drugs that specifically target the inverted methylation pattern of cancerous cells include the DNA methyltransferase inhibitors azacitidine and decitabine. These hypomethylating agents are used to treat myelodysplastic syndrome, a blood cancer produced by abnormal bone marrow stem cells. These agents inhibit all three types of active DNA methyltransferases, and had been thought to be highly toxic, but proved to be effective when used in low dosage, reducing progression of myelodysplastic syndrome to leukemia.

Histone deacetylase (HDAC) inhibitors show efficacy in treatment of T cell lymphoma. two HDAC inhibitors, vorinostat and romidepsin, have been approved by the Food and Drug Administration. However, since these HDAC inhibitors alter the acetylation state of many proteins in addition to the histone of interest, knowledge of the underlying mechanism at the molecular level of patient response is required to enhance the efficiency of using such inhibitors as treatment. Treatment with HDAC inhibitors has been found to promote gene reactivation after DNA methyl-transferases inhibitors have repressed transcription. Panobinostat is approved for certain situations in myeloma.

Other pharmaceutical targets in research are histone lysine methyltransferases (KMT) and protein arginine methyltransferases (PRMT). Preclinical study has suggested that lunasin may have potentially beneficial epigenetic effects.

DNA methylation in cancer

From Wikipedia, the free encyclopedia

DNA methylation in cancer plays a variety of roles, helping to change the healthy regulation of gene expression to a disease pattern.

All mammalian cells descended from a fertilized egg (a zygote) share a common DNA sequence (except for new mutations in some lineages). However, during development and formation of different tissues epigenetic factors change. The changes include histone modifications, CpG island methylations and chromatin reorganizations which can cause the stable silencing or activation of particular genes. Once differentiated tissues are formed, CpG island methylation is generally stably inherited from one cell division to the next through the DNA methylation maintenance machinery.

In cancer, a number of mutational changes are found in protein coding genes. Colorectal cancers typically have 3 to 6 driver mutations and 33 to 66 hitchhiker or passenger mutations that silence protein expression in the genes affected. However, transcriptional silencing may be more important than mutation in causing gene silencing in progression to cancer. In colorectal cancers about 600 to 800 genes are transcriptionally silenced, compared to adjacent normal-appearing tissues, by CpG island methylation. Transcriptional repression in cancer can also occur by other epigenetic mechanisms, such as altered expression of microRNAs.

CpG islands are frequent control elements

CpG islands are commonly 200 to 2000 base pairs long, have a C:G base pair content >50%, and have frequent 5' → 3' CpG sequences. About 70% of human promoters located near the transcription start site of a gene contain a CpG island.

Promoters located at a distance from the transcription start site of a gene also frequently contain CpG islands. The promoter of the DNA repair gene ERCC1, for instance, was identified and located about 5,400 nucleotides upstream of its coding region. CpG islands also occur frequently in promoters for functional noncoding RNAs such as microRNAs and Long non-coding RNAs (lncRNAs).

Methylation of CpG islands in promoters stably silences genes

Genes can be silenced by multiple methylation of CpG sites in the CpG islands of their promoters. Even if silencing of a gene is initiated by another mechanism, this often is followed by methylation of CpG sites in the promoter CpG island to stabilize the silencing of the gene. On the other hand, hypomethylation of CpG islands in promoters can result in gene over-expression.

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 instance, in colon tumors compared to adjacent normal-appearing colonic mucosa, about 600 to 800 heavily methylated CpG islands occur in promoters of genes in the tumors while these CpG islands are not methylated in the adjacent mucosa. In contrast, as Vogelstein et al. point out, in a colorectal cancer there are typically only about 3 to 6 driver mutations and 33 to 66 hitchhiker or passenger mutations.

DNA repair gene silencing in cancer

In sporadic cancers, a DNA repair deficiency is occasionally found to be due to a mutation in a DNA repair gene. However, much more frequently, reduced or absent expression of a DNA repair gene in cancer is due to methylation of its promoter. For example, of 113 colorectal cancers examined, only four had a missense mutation in the DNA repair gene MGMT, while the majority had reduced MGMT expression due to methylation of the MGMT promoter region. Similarly, among 119 cases of mismatch repair-deficient colorectal cancers that lacked DNA repair gene PMS2 expression, 6 had a mutation in the PMS2 gene, while for 103 PMS2 was deficient because its pairing partner MLH1 was repressed due to promoter methylation (PMS2 protein is unstable in the absence of MLH1). In the remaining 10 cases, loss of PMS2 expression was likely due to epigenetic overexpression of the microRNA, miR-155, which down-regulates MLH1.

Frequency of hypermethylation of DNA repair genes in cancer

Twenty-two DNA repair genes with hypermethylated promoters, and reduced or absent expression, were found to occur among 17 types of cancer, as listed in two review articles. As listed in one of the reviews, promoter hypermethylation of MGMT occurs frequently in a number of cancers including 93% of bladder cancers, 88% of stomach cancers, 74% of thyroid cancers, 40%-90% of colorectal cancers and 50% of brain cancers. That review also indicated promoter hypermethylation of LIG4, NEIL1, ATM, MLH1 or FANCB occurs at frequencies between 33% to 82% in one or more of head and neck cancers, non-small-cell lung cancers or non-small-cell lung cancer squamous cell carcinomas. The article Werner syndrome ATP-dependent helicase indicates the DNA repair gene WRN has a promoter that is frequently hypermethylated in a number of cancers, with hypermethylation occurring in 11% to 38% of colorectal, head and neck, stomach, prostate, breast, thyroid, non-Hodgkin lymphoma, chondrosarcoma and osteosarcoma cancers.

Likely role of hypermethylation of DNA repair genes in cancer

As discussed by Jin and Roberston in their review, silencing of a DNA repair gene by hypermethylation may be a very early step in progression to cancer. Such silencing is proposed to act similarly to a germ-line mutation in a DNA repair gene, and predisposes the cell and its descendants to progression to cancer. Another review also indicated an early role for hypermethylation of DNA repair genes in cancer. If a gene necessary for DNA repair is hypermethylated, resulting in deficient DNA repair, DNA damages will accumulate. Increased DNA damage tends to cause increased errors during DNA synthesis, leading to mutations that can give rise to cancer. 

If hypermethylation of a DNA repair gene is an early step in carcinogenesis, then it may also occur in the normal-appearing tissues surrounding the cancer from which the cancer arose (the field defect). See the table below. 

Frequencies of hypermethylated promoters in DNA repair genes in sporadic cancers and in adjacent field defects
Cancer Gene Frequency in Cancer Frequency in Field Defect
Colorectal MGMT 55% 54%
Colorectal MSH2 13% 5%
Colorectal WRN 29% 13%
Head and Neck MGMT 54% 38%
Head and Neck MLH1 33% 25%
Non-small cell lung cancer ATM 69% 59%
Non-small cell lung cancer MLH1 69% 72%
Stomach MGMT 88% 78%
Stomach MLH1 73% 20%
Esophagus MLH1 77%-100% 23%-79%

While DNA damages may give rise to mutations through error prone translesion synthesis, DNA damages can also give rise to epigenetic alterations during faulty DNA repair processes. The DNA damages that accumulate due to hypermethylation of the promoters of DNA repair genes can be a source of the increased epigenetic alterations found in many genes in cancers. 

In an early study, looking at a limited set of transcriptional promoters, Fernandez et al. examined the DNA methylation profiles of 855 primary tumors. Comparing each tumor type with its corresponding normal tissue, 729 CpG island sites (55% of the 1322 CpG island sites evaluated) showed differential DNA methylation. Of these sites, 496 were hypermethylated (repressed) and 233 were hypomethylated (activated). Thus, there is a high level of promoter methylation alterations in tumors. Some of these alterations may contribute to cancer progression.

DNA methylation of microRNAs in cancer

In mammals, microRNAs (miRNAs) regulate the transcriptional activity of about 60% of protein-encoding genes. Individual miRNAs can each target, and repress transcription of, on average, roughly 200 messenger RNAs of protein coding genes. The promoters of about one third of the 167 miRNAs evaluated by Vrba et al. in normal breast tissues were differentially hyper/hypo-methylated in breast cancers. A more recent study pointed out that the 167 miRNAs evaluated by Vrba et al. were only 10% of the miRNAs found expressed in breast tissues. This later study found that 58% of the miRNAs in breast tissue had differentially methylated regions in their promoters in breast cancers, including 278 hypermethylated miRNAs and 802 hypomethylated miRNAs.

One miRNA that is over-expressed about 100-fold in breast cancers is miR-182. MiR-182 targets the BRCA1 messenger RNA and may be a major cause of reduced BRCA1 protein expression in many breast cancers.

microRNAs that control DNA methyltransferase genes in cancer

Some miRNAs target the messenger RNAs for DNA methyltransferase genes DNMT1, DNMT3A and DNMT3B, whose gene products are needed for initiating and stabilizing promoter methylations. As summarized in three reviews, miRNAs miR-29a, miR-29b and miR-29c target DNMT3A and DNMT3B; miR-148a and miR-148b target DNMT3B; and miR-152 and miR-301 target DNMT1. In addition, miR-34b targets DNMT1 and the promoter of miR-34b itself is hypermethylated and under-expressed in the majority of prostate cancers. When expression of these microRNAs is altered, they may also be a source of the hyper/hypo-methylation of the promoters of protein-coding genes in cancers.

Genome instability

From Wikipedia, the free encyclopedia

Genome instability (also genetic instability or genomic instability) refers to a high frequency of mutations within the genome of a cellular lineage. These mutations can include changes in nucleic acid sequences, chromosomal rearrangements or aneuploidy. Genome instability does occur in bacteria. In multicellular organisms genome instability is central to carcinogenesis, and in humans it is also a factor in some neurodegenerative diseases such as amyotrophic lateral sclerosis or the neuromuscular disease myotonic dystrophy

The sources of genome instability have only recently begun to be elucidated. A high frequency of externally caused DNA damage can be one source of genome instability since DNA damages can cause inaccurate translesion synthesis past the damages or errors in repair, leading to mutation. Another source of genome instability may be epigenetic or mutational reductions in expression of DNA repair genes. Because endogenous (metabolically-caused) DNA damage is very frequent, occurring on average more than 60,000 times a day in the genomes of human cells, any reduced DNA repair is likely an important source of genome instability.

The usual genome situation

Usually, all cells in an individual in a given species (plant or animal) show a constant number of chromosomes, which constitute what is known as the karyotype defining this species (see also List of number of chromosomes of various organisms), although some species present a very high karyotypic variability. In humans, mutations that would change an amino acid within the protein coding region of the genome occur at an average of only 0.35 per generation (less than one mutated protein per generation).

Sometimes, in a species with a stable karyotype, random variations that modify the normal number of chromosomes may be observed. In other cases, there are structural alterations (chromosomal translocations, deletions ...) that modify the standard chromosomal complement. In these cases, it is indicated that the affected organism presents genome instability (also genetic instability, or even chromosomic instability). The process of genome instability often leads to a situation of aneuploidy, in which the cells present a chromosomic number that is either higher or lower than the normal complement for the species.

Causes of genome instability

DNA Replication Defects

In the cell cycle, DNA is usually most vulnerable during replication. The replisome must be able to navigate obstacles such as tightly wound chromatin with bound proteins, single and double stranded breaks which can lead to the stalling of the replication fork. Each protein or enzyme in the replisome must perform its function well to result in a perfect copy of DNA. Mutations of proteins such as DNA polymerase, ligase, can lead to impairment of replication and lead to spontaneous chromosomal exchanges. Proteins such as Tel1, Mec1 (ATR, ATM in humans) can detect single and double-stranded breaks and recruit factors such as Rmr3 helicase to stabilize the replication fork in order to prevent its collapse. Mutations in Tel1, Mec1, and Rmr3 helicase result in a significant increase of chromosomal recombination. ATR responds specifically to stalled replication forks and single-stranded breaks resulting from UV damage while ATM responds directly to double-stranded breaks. These proteins also prevent progression into mitosis by inhibiting the firing of late replication origins until the DNA breaks are fixed by phosphorylating CHK1, CHK2 which results in a signaling cascade arresting the cell in S-phase. For single stranded breaks, replication occurs until the location of the break, then the other strand is nicked to form a double stranded break, which can then be repaired by Break Induced Replication or homologous recombination using the sister chromatid as an error-free template. In addition to S-phase checkpoints, G1 and G2 checkpoints exist to check for transient DNA damage which could be caused by mutagens such as UV damage. An example is the Saccharomyces pombe gene rad9 which arrests the cells in late S/G2 phase in the presence of DNA damage caused by radiation. The yeast cells with defective rad9 failed to arrest following radiation, continued cell division and died rapidly while the cells with wild-type rad9 successfully arrested in late S/G2 phase and remained viable. The cells that arrested were able to survive due to the increased time in S/G2 phase allowing for DNA repair enzymes to function fully.

Fragile Sites

There are hotspots in the genome where DNA sequences are prone to gaps and breaks after inhibition of DNA synthesis such as in the aforementioned checkpoint arrest. These sites are called fragile sites, and can occur commonly as naturally present in most mammalian genomes or occur rarely as a result of mutations, such as DNA-repeat expansion. Rare fragile sites can lead to genetic disease such as fragile X mental retardation syndrome, myotonic dystrophy, Friedrich’s ataxia, and Huntington’s disease, most of which are caused by expansion of repeats at the DNA, RNA, or protein level. Although, seemingly harmful, these common fragile sites are conserved all the way to yeast and bacteria. These ubiquitous sites are characterized by trinucleotide repeats, most commonly CGG, CAG, GAA, and GCN. These trinucleotide repeats can form into hairpins, leading to difficulty of replication. Under replication stress, such as defective machinery or further DNA damage, DNA breaks and gaps can form at these fragile sites. Using a sister chromatid as repair is not a fool-proof backup as the surrounding DNA information of the n and n+1 repeat is virtually the same, leading to copy number variation. For example, the 16th copy of CGG might be mapped to the 13th copy of CGG in the sister chromatid since the surrounding DNA is both CGGCGGCGG…, leading to 3 extra copies of CGG in the final DNA sequence.

Transcription-associated instability

In both E. coli and Saccromyces pombe, transcription sites tend to have higher recombination and mutation rates. The coding or non-transcribed strand accumulates more mutations than the template strand. This is due to the fact that the coding strand is single-stranded during transcription, which is chemically more unstable than double-stranded DNA. During elongation of transcription, supercoiling can occur behind an elongating RNA polymerase, leading to single-stranded breaks. When the coding strand is single-stranded, it can also hybridize with itself, creating DNA secondary structures that can compromise replication. In E. coli, when attempting to transcribe GAA triplets such as those found in Friedrich’s ataxia, the resulting RNA and template strand can form mismatched loops between different repeats, leading the complementary segment in the coding-strand available to form its own loops which impede replication. Furthermore, replication of DNA and transcription of DNA are not temporally independent; they can occur at the same time and lead to collisions between the replication fork and RNA polymerase complex. In S. cerevisiae, Rrm3 helicase is found at highly transcribed genes in the yeast genome, which is recruited to stabilize a stalling replication fork as described above. This suggests that transcription is an obstacle to replication, which can lead to increased stress in the chromatin spanning the short distance between the unwound replication fork and transcription start site, potentially causing single-stranded DNA breaks. In yeast, proteins act as barriers at the 3’ of the transcription unit to prevent further travel of the DNA replication fork.

Increase Genetic Variability

In some portions of the genome, variability is essential to survival. One such locale is the Ig genes. In a pre-B cell, the region consists of all V, D, and J segments. During development of the B cell, a specific V, D, and J segment is chosen to be spliced together to form the final gene, which is catalyzed by RAG1 and RAG2 recombinases. Activation-Induced Cytidine Deaminase (AID) then converts cytidine into uracil. Uracil normally does not exist in DNA, and thus the base is excised and the nick is converted into a double-stranded break which is repaired by non-homologous end joining (NHEJ). This procedure is very error-prone and leads to somatic hypermutation. This genomic instability is crucial in ensuring mammalian survival against infection. V, D, J recombination can ensure millions of unique B-cell receptors; however, random repair by NHEJ introduces variation which can create a receptor that can bind with higher affinity to antigens.

In neuronal and neuromuscular disease

Of about 200 neurological and neuromuscular disorders, 15 have a clear link to an inherited or acquired defect in one of the DNA repair pathways or excessive genotoxic oxidative stress. Five of them (xeroderma pigmentosum, Cockayne's syndrome, trichothiodystrophy, Down's syndrome, and triple-A syndrome) have a defect in the DNA nucleotide excision repair pathway. Six (spinocerebellar ataxia with axonal neuropathy-1, Huntington's disease, Alzheimer's disease, Parkinson's disease, Down's syndrome and amyotrophic lateral sclerosis) seem to result from increased oxidative stress, and the inability of the base excision repair pathway to handle the damage to DNA that this causes. Four of them (Huntington's disease, various spinocerebellar ataxias, Friedreich’s ataxia and myotonic dystrophy types 1 and 2) often have an unusual expansion of repeat sequences in DNA, likely attributable to genome instability. Four (ataxia-telangiectasia, ataxia-telangiectasia-like disorder, Nijmegen breakage syndrome and Alzheimer's disease) are defective in genes involved in repairing DNA double-strand breaks. Overall, it seems that oxidative stress is a major cause of genomic instability in the brain. A particular neurological disease arises when a pathway that normally prevents oxidative stress is deficient, or a DNA repair pathway that normally repairs damage caused by oxidative stress is deficient.

In cancer

In cancer, genome instability can occur prior to or as a consequence of transformation. Genome instability can refer to the accumulation of extra copies of DNA or chromosomes, chromosomal translocations, chromosomal inversions, chromosome deletions, single-strand breaks in DNA, double-strand breaks in DNA, the intercalation of foreign substances into the DNA double helix, or any abnormal changes in DNA tertiary structure that can cause either the loss of DNA, or the misexpression of genes. Situations of genome instability (as well as aneuploidy) are common in cancer cells, and they are considered a "hallmark" for these cells. The unpredictable nature of these events are also a main contributor to the heterogeneity observed among tumour cells. 

It is currently accepted that sporadic tumors (non-familial ones) are originated due to the accumulation of several genetic errors. An average cancer of the breast or colon can have about 60 to 70 protein altering mutations, of which about 3 or 4 may be "driver" mutations, and the remaining ones may be "passenger" mutations Any genetic or epigenetic lesion increasing the mutation rate will have as a consequence an increase in the acquisition of new mutations, increasing then the probability to develop a tumor. During the process of tumorogenesis, it is known that diploid cells acquire mutations in genes responsible for maintaining genome integrity (caretaker genes), as well as in genes that are directly controlling cellular proliferation (gatekeeper genes). Genetic instability can originate due to deficiencies in DNA repair, or due to loss or gain of chromosomes, or due to large scale chromosomal reorganizations. Losing genetic stability will favour tumor development, because it favours the generation of mutants that can be selected by the environment.

The tumor microenvironment has an inhibitory effect on DNA repair pathways contributing to genomic instability, which promotes tumor survival, proliferation, and malignant transformation.

Low frequency of mutations without cancer

The protein coding regions of the human genome, collectively called the exome, constitutes only 1.5% of the total genome. As pointed out above, ordinarily there are only an average of 0.35 mutations in the exome per generation (parent to child) in humans. In the entire genome (including non-protein coding regions) there are only about 70 new mutations per generation in humans.

Cause of mutations in cancer

The likely major underlying cause of mutations in cancer is DNA damage. For example, in the case of lung cancer, DNA damage is caused by agents in exogenous genotoxic tobacco smoke (e.g. acrolein, formaldehyde, acrylonitrile, 1,3-butadiene, acetaldehyde, ethylene oxide and isoprene). Endogenous (metabolically-caused) DNA damage is also very frequent, occurring on average more than 60,000 times a day in the genomes of human cells (see DNA damage (naturally occurring)). Externally and endogenously caused damages may be converted into mutations by inaccurate translesion synthesis or inaccurate DNA repair (e.g. by non-homologous end joining). In addition, DNA damages can also give rise to epigenetic alterations during DNA repair. Both mutations and epigenetic alterations (epimutations) can contribute to progression to cancer.

Very frequent mutations in cancer

As noted above, about 3 or 4 driver mutations and 60 passenger mutations occur in the exome (protein coding region) of a cancer. However, a much larger number of mutations occur in the non-protein-coding regions of DNA. The average number of DNA sequence mutations in the entire genome of a breast cancer tissue sample is about 20,000. In an average melanoma tissue sample (where melanomas have a higher exome mutation frequency) the total number of DNA sequence mutations is about 80,000.

Cause of high frequency of mutations in cancer

The high frequency of mutations in the total genome within cancers suggests that, often, an early carcinogenic alteration may be a deficiency in DNA repair. Mutation rates substantially increase (sometimes by 100-fold) in cells defective in DNA mismatch repair or in homologous recombinational DNA repair. Also, chromosomal rearrangements and aneuploidy increase in humans defective in DNA repair gene BLM.

A deficiency in DNA repair, itself, can allow DNA damages to accumulate, and error-prone translesion synthesis past some of those damages may give rise to mutations. In addition, faulty repair of these accumulated DNA damages may give rise to epigenetic alterations or epimutations. While a mutation or epimutation in a DNA repair gene, itself, would not confer a selective advantage, such a repair defect may be carried along as a passenger in a cell when the cell acquires an additional mutation/epimutation that does provide a proliferative advantage. Such cells, with both proliferative advantages and one or more DNA repair defects (causing a very high mutation rate), likely give rise to the 20,000 to 80,000 total genome mutations frequently seen in cancers.

DNA repair deficiency in cancer

In somatic cells, deficiencies in DNA repair sometimes arise by mutations in DNA repair genes, but much more often are due to epigenetic reductions in expression of DNA repair genes. Thus, in a sequence of 113 colorectal cancers, only four had somatic missense mutations in the DNA repair gene MGMT, while the majority of these cancers had reduced MGMT expression due to methylation of the MGMT promoter region. Five reports, listed in the article Epigenetics (see section "DNA repair epigenetics in cancer") presented evidence that between 40% and 90% of colorectal cancers have reduced MGMT expression due to methylation of the MGMT promoter region. 

Similarly, for 119 cases of colorectal cancers classified as mismatch repair deficient and lacking DNA repair gene PMS2 expression, Pms2 was deficient in 6 due to mutations in the PMS2 gene, while in 103 cases PMS2 expression was deficient because its pairing partner MLH1 was repressed due to promoter methylation (PMS2 protein is unstable in the absence of MLH1). The other 10 cases of loss of PMS2 expression were likely due to epigenetic overexpression of the microRNA, miR-155, which down-regulates MLH1.

In cancer epigenetics (see section Frequencies of epimutations in DNA repair genes), there is a partial listing of epigenetic deficiencies found in DNA repair genes in sporadic cancers. These include frequencies of between 13–100% of epigenetic defects in genes BRCA1, WRN, FANCB, FANCF, MGMT, MLH1, MSH2, MSH4, ERCC1, XPF, NEIL1 and ATM located in cancers including breast, ovarian, colorectal and head and neck. Two or three epigenetic deficiencies in expression of ERCC1, XPF and/or PMS2 were found to occur simultaneously in the majority of the 49 colon cancers evaluated. Some of these DNA repair deficiencies can be caused by epimutations in microRNAs as summarized in the MicroRNA article section titled miRNA, DNA repair and cancer.

Lymphomas as a consequence of genome instability

Cancers usually result from disruption of a tumor repressor or dysregulation of an oncogene. Knowing that B-cells experience DNA breaks through development can give insight to the genome of lymphomas. Many types of lymphoma are caused by chromosomal translocation, which can arise from breaks in DNA leading to incorrect joining. In Burkitt’s lymphoma, c-myc, an oncogene encoding a transcription factor, is translocated after the promoter of the immunoglobulin gene, leading dysregulation of c-myc transcription. Since immunoglobulins are essential to a lymphocyte and highly expressed to increase detection of antigens, c-myc is then also highly expressed leading to transcription of its targets which are involved in cell proliferation. Mantle cell lymphoma is characterized by fusion of cyclin D1 to the immunoglobulin locus. Cyclin D1 inhibits Rb, a tumor suppressor, leading to tumorigenesis. Follicular lymphoma results from the translocation of immunoglobulin promoter to the Bcl-2 gene, giving rise to large amounts of Bcl-2 protein which inhibits apoptosis. DNA-damaged B-cells no longer undergo apoptosis leading to further mutations which could affect driver genes leading to tumorigenesis. The location of translocation in the oncogene shares structural properties of the target regions of AID, suggesting that the oncogene was a potential target of AID, leading to a double-stranded break that was translocated to the immunoglobulin gene locus through NHEJ repair.

Welfare economics

From Wikipedia, the free encyclopedia https://en.wikipedia.org/wiki/Welfare_economics   Welfare economics is a...