Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 2514 word(s) 2514 2021-10-11 05:29:36 |
2 References added. Meta information modification 2514 2021-10-11 18:25:22 | |
3 format correct Meta information modification 2514 2021-10-12 04:03:55 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Maurya, S. Enhancers in Development and Diseases. Encyclopedia. Available online: https://encyclopedia.pub/entry/14957 (accessed on 25 April 2024).
Maurya S. Enhancers in Development and Diseases. Encyclopedia. Available at: https://encyclopedia.pub/entry/14957. Accessed April 25, 2024.
Maurya, Shailendra. "Enhancers in Development and Diseases" Encyclopedia, https://encyclopedia.pub/entry/14957 (accessed April 25, 2024).
Maurya, S. (2021, October 11). Enhancers in Development and Diseases. In Encyclopedia. https://encyclopedia.pub/entry/14957
Maurya, Shailendra. "Enhancers in Development and Diseases." Encyclopedia. Web. 11 October, 2021.
Enhancers in Development and Diseases
Edit

Enhancers are cis-regulatory elements containing short DNA sequences that serve as binding sites for pioneer/regulatory transcription factors, thus orchestrating the regulation of genes critical for lineage determination. The activity of enhancer elements is believed to be determined by transcription factor binding, thus determining the cell state identity during development. Precise spatio-temporal control of the transcriptome during lineage specification requires the coordinated binding of lineage-specific transcription factors to enhancers. Thus, enhancers are the primary determinants of cell identity. Numerous studies have explored the role and mechanism of enhancers during development and disease, and various basic questions related to the functions and mechanisms of enhancers have not yet been fully answered.

enhancer poised cis-regulatory lineage spatio-temporal

1. Introduction

The term “enhancer” was first coined based on studies of Simian virus 40 (SV40) when, in 1981, Banerji et al. observed for the first time that a viral DNA element from SV40 had the ability to enhance activity towards a T-antigen or β-globin reporter in mammalian cells [1]. Further research has explored endogenous sequences with similar functions in the immunoglobulin heavy chain locus. These preliminary studies have established that enhancers function as short DNA elements that trigger a gene’s transcription from a long distance in an orientation-independent manner. Preliminary studies have determined the exact mechanism of how distal regulatory elements regulate gene transcription through distal enhancers [2]. Enhancers are cis-regulatory elements that carry epigenetic information in DNA sequences through specific histone modifications [3]. Studies assessing the characteristics of enhancers have reported that they can function independently of the orientation and distance to their cognate target genes, at distances sometimes of several hundred kilobases or megabases [2][4][5]. Enhancers can be identified and characterized by various factors, including histone modifications, their transcription into non-coding RNAs, and their epigenetic features [6]. The prominent feature of enhancers is their ability to serve as a docking platform for transcription factor binding, where developmental signaling (intrinsic or extrinsic) cues are interpreted in a highly context-specific manner [7]. The signatures of these enhancers are highly cell type-specific; such cell type-specific use of the epigenetic information has demonstrated the combinatorial function of transcription factors in maintaining cell identity and lineage determination [8]. The cell type-specific enhancer pattern provides a unique cis-regulatory platform, in which transcription factors are activated by developmental cues and modify the transcriptome [9][10][11]. This model revealed that every developmental stage has a cell type-specific transcription factor, which functions by cell type-specific enhancer signatures [12]. Enhancers are actuated in a stage-specific manner, correlated with cell type-specific histone modifications. This combinatorial mechanism of transcription factor binding on cell type-specific enhancers results in the so-called enhancer signature, which serves as a readout to define enhancers in a cell type-specific manner at a global scale [13]. Precise spatial and temporal control of gene expression and the correct interpretation of the enhancer signature are crucial in the development process. Any alterations in the enhancer signature can modify the gene expression pattern and ethe capacity of the enhancer to respond to developmental signals, narrowing cell differentiation and affecting correct lineage formation. Heinz et al. (2010) have shown that lineage-determining transcription factors bind to a genomic region in a cell-specific manner. They showed the genome-wide locations of PU.1 binding patterns in macrophages, B-cells, and diffuse B-cell progenitors [14]. Similarly, Xu et al. (2012) analyzed chromatin state maps, transcription factor occupancy rates, and gene expression profiles during the development of human erythroid cells at the fetal and adult stages, and carried out a comparative analysis to determine the specific procedures of the development stage [15]. Here, it is also important to discuss the study of Choukrallah et al. (2015), who found that the enhancer landscape is dynamically reshaped in each differentiation step. Interesting changes include creating new enhancers and the closing and re-opening of the pre-existing enhancer landscape [16]. The authors also reported that the regulatory signatures of two related types of myeloid leukemia expression fusion proteins (RUNX1-ETO and RUNX1-EV1) display a distinct binding pattern and interact with different transcription factors to impact the epigenome [17]. A study has also reported that MLL-Af9 and MLL-AF4 oncofusion proteins showed distinct binding patterns in the enhancer region and targeted the RUNX1 program in 11q23 acute myeloid leukemia [18]. Enhancers are essential for normal functioning, and the loss of enhancer elements can cause abnormalities; for example, Groschel et al. (2014) showed that the removal of the distal enhancer essential for the GATA2 gene resulted in insufficient Functional GATA2 haploids, which only reduced the expression of the remaining normal alleles [19]. This review focuses on a concise and brief overview of the roles of enhancers in development and disease. We attempt to discuss how enhancers are activated and coordinated with transcription factors, as well as the roles of enhancers in mammalian development. This may comprise the first attempt to compile recently published research on development and disease, focusing on enhancers.

2. Features and Types of Enhancers

Enhancers have been identified in the form of various regulatory domains, primed enhancers, active enhancers, and poised enhancers. Each type of enhancer signature has specific histone modification patterns and can be easily identified by these signatures [13][20][21]. Primed enhancers can be identified by only histone H3K4 mono-methylation, while active enhancers signatures are identified by H3K4 mono-methylation and H3K27ac. Finally, the signatures of poised enhancers are marked with H3K4Me1 with H3K27me3, but not H3K27ac [13][20][21]. Active enhancers are linked to expressed genes, while poised enhancers are always associated to developmental genes, which are inactive in embryonic stem cells or precursor cells and become expressed during different differentiation stages [22]. During differentiation, the poised enhancer’s signature successfully loses the repressive H3K27me3 histone mark, acquires H3K27ac marks, and becomes active. Enhancers are, thus, subject to dynamic change functions, as an on/off switch to tune the target gene expression and changing the cell state from undifferentiated to differentiated phenotype [23]. Hence, the signatures of Poised Enhancers comprise a small set of regulatory signatures in embryonic stem cells that facilitate their timely and stage-specific function, once the correct differentiation signals become available [20]. One term also uses super-enhancers (SEs), which are described as large clusters of active enhancers with robust enrichment for binding transcriptional coactivators [24][25]. These features have also been called stretch enhancers [26], multiple enhancers [27], and enhancer clusters [28], which are similar but not identical between studies (although many of these features overlap). SEs regulate master regulators of pluripotency, such as OCT4, SOX2, and NANOG [24]. It has been reported that the SE signature is often enriched near the oncogene in tumor cells, while an enrichment GWAS has identified SNPs normally associated with several common diseases [29][30].

3. Role of Enhancers in Disease Development

The genome-wide sequencing approach has revealed that enhancers are prime targets for genetic or epigenetic alterations that lead to carcinogenesis [5]. The main characteristics of the enhancer signature remain constant for every type of cell (e.g., normal vs. cancer/tumor cells). The difference is that the function of the enhancer signatures differs between normal and cancer cell types; that is, the enhancer’s functional output in tumor/cancer cells differs from that in normal cells [31]. Careful examination of enhancer signatures between normal cells and their counterpart in cancer cells has revealed that cancer cells tend to lose enhancers at positions near cell fate-specifying genes and gain enhancer signatures near growth-associated genes [32][33][34]. Furthermore, it has again been confirmed, after comparing normal colon epithelial crypts and colon cancer lines, that thousands of differentially enriched primed enhancer marks (H3K4Me1) were enriched in cancerous cells known as variant enhancer loci (Veli), and these gained mono-methylation sites were not found in normal counterpart cells. In contrast, the lost sites were relatively specific to crypt cells, indicating that the colon cancer cells acquired a more differentiated cell state [33]. In the post-GWAS era, there is a lot of conceivable evidence that cancer predisposition genomic variants present in the non-coding genomic region lie within these distal regulatory elements [35][36]. This finding is based on the overlap between SNPs associated with diseases and genomic signatures, revealing that SNPs associated with risk phenotypes are frequently enriched in expression quantitative traits and open chromatin regions [37][38]. In cancer cells, risk variants lead to the dysregulation of enhancers, disrupting the fine-tuned target expression of their associated genes and producing a pathological state which leads to abnormal growth. For example, pancreatic-specific differential open regions enriched for non-coding variants (SNPs) have been linked to pancreatic disorders [39]. Correspondingly, monocyte-/macrophage-specific enhancer elements enriched with SNPs associated with ulcerative colitis, celiac disease, Crohn’s disease, or systemic lupus erythematosus have been reported [40]. Studies are also available in which single base-pair point mutations in the distal enhancer of SCNA and PTF1A have been shown to cause sporadic Parkinson’s disease and pancreatic agenesis [41][42]. Altered super-enhancer activity has also been reported in many complex human diseases, such as Alzheimer’s, Type 1 diabetes, and autoimmune disorders [25][28][43][44]. Furthermore, lost super-enhancers and enhancers acquired somatically have been reported to be associated with numerous cancers [25][43]. Meanwhile, changes in histone writers, erasers, and altered activity of DNA modifiers such as methyl and acetyltransferases also reshape the chromatin landscape, which finally leads to abnormal growth and development. Based on known enhancer signatures, several studies (Table 1) have explored the alterations in enhancer activity and histone modification patterns which may be correlated with disease development. Evidence for enhancer alterations in cancer, coming from various studies, is summarized in Table 1.
Table 1. Studies investigating the roles of Enhancers in disease development.
Number Title Author Reference
1 Translocation of the c-myc gene into the immunoglobulin heavy chain locus in human Burkitt lymphoma and murine plasmacytoma cells Taub et al., 1982 [45]
2 A long-range Shh enhancer regulates expression in the developing limb and fin and is associated with preaxial polydactyly Lettice, L.A. et al., 2003 [46]
3 Genomic deletion of a long-range bone enhancer misregulates in Van Buchem disease Loots, G.G. et al., 2005 [47]
4 A common sex dependent mutations in a RET enhancer underlies Hirschsprung disease risk Emison, E.S. et al., 2005 [48]
5 Disruption of an AP2-alpha binding site in an IRF6 enhancer is associated with cleft lip Rahimov et al., 2008 [49]
6 Functional enhancers at the gene-poor 8q24 cancer linked locus Jia, L. et al., 2009 [50]
7 The 8q24 cancer risk variant rs6983267 shows long-range interaction with MYC in colorectal cancer Pomeratz, M.M. et al., 2009 [51]
8 The common colorectal cancer predisposition SNP rs6983267 at chromosome 8q24 confers potential to enhanced WNT signaling Tuupanen, S. et al., 2009 [52]
9 Long-range enhancers on 8q24 regulate c-Myc Sotelo et al., 2010 [53]
10 An 8q24 gene desert variant associated with prostate cancer risk confers differential in vivo activity to a MYC enhancer. Wasserman, N.F. et al., 2010 [54]
11 Enhancer-adoption as a mechanism of human developmental disease Lettice, L.A. et al., 2011 [55]
12 Systematic localization of common disease associated variation in regulatory DNA Maurano et al., 2012 [36]
13 Epigenomic enhancer profiling defines a signature of colon cancer Akhtar-Zaidi, B. et al., 2012 [33]
14 Mice lacking a Myc enhancer that includes human SNP rs6983267 are resistant to intestinal tumors Sur et al., 2012 [56]
15 A novel 13 base pair insertion in the sonic hedgehog ZRS limb enhancer (LMBR1) causes preaxial polydactyly with triphalangeal thumb Laurell, T. et al., 2012 [57]
16 Regulatory variation in a TBX5 enhancer leads to isolated congenital heart diseases Smemo, S. et al., 2012 [58]
17 DNA methylation of transcriptional enhancers and cancer predisposition Aran and Hallman et al., 2013 [59]
18 Discovery and characterization of super-enhancer associated dependencies in diffuse large B cell lymphoma Chapuy, B. et al., 2013 [60]
19 Chromatin stretch enhancer states drive cell specific gene regulation and harbor human disease risk variants Parker, S.C. et al., 2013 [26]
20 Role of SWI/SNF in acute leukemia maintenance and enhancer-mediated Myc regulation Shi, J. et al., 2013 [61]
21 Selective inhibition of tumor oncogenes by disruption of super-enhancers Loven, J. et al., 2013 [29]
22 An erythroid enhancer of BCL11A subject to genetic variation determines fetal hemoglobin level Bauer, D.E. et al., 2013 [62]
23 Disruption of autoregulatory feedback by a mutation in a remote, ultraconserved PAX6 enhancer causes aniridia Bhatia, S. et al., 2013 [63]
24 Genome-wide analysis of noncoding regulatory mutations in cancer Weinhold, N. et al., 2014 [64]
25 A NOTCH driven MYC enhancer promotes T cell development, transformation and acute lymphoblastic leukemia Herranz, D. et al., 2014 [65]
26 Combinatorial effects of multiple enhancer variants linkage disequilibrium dictate levels of gene expression to confer susceptibility to common traits Corradin, O. et al., 2014 [27]
27 Enhancer hijacking activates GFI1 family oncogenes in medulloblastoma Northcott, P.A. et al., 2014 [66]
28 Microduplications encompassing the sonic hedgehog limb enhancer ZRS are associated with Hass-type polysyndactyly and Laurin Sandrow syndrome Lohan, S. et al., 2014 [67]
29 Epigenomic analysis of primary human T cells reveals enhancers associated with TH2 memory cell differentiation and asthma susceptibility Seumois, G. et al., 2014 [68]
30 Oncogenic regulation. An oncogenic Super enhancer formed through somatic mutation of a noncoding intergenic element. Mansour, M.R. et al., 2014 [69]
31 A Sox2 distal enhancer cluster regulates embryonic stem cell differentiation potential Zhou, H.Y. et al., 2014 [70]
32 Multiple functional risk variants in a SMAD7 enhancer implicate a colorectal cancer risk haplotype Fortini, B.L. et al., 2014 [71]
33 A remote GATA2 hematopoietic enhancer drives leukemogenesis in inv(3) (q21;q26) by activating EVI1 expression Yamazaki et al., 2014 [72]
34 Long range enhancer activity determines Myc sensitivity to Notch inhibitors in T cell leukemia Yashiro-Ohtani et al., 2014 [73]
35 Recessive mutations in a distal PTF1A enhancer cause isolated pancreatic agenesis Weedon et al., 2014 [42]
36 A single oncogenic enhancer rearrangement causes concomitant EV1 and GATA2 deregulation in leukemia Groschel et al., 2014 [19]
37 Genetic predisposition to neuroblastoma mediated by a LMO1 super enhancer polymorphisms Oldridge, D.A. et al., 2015 [74]
38 7q21.3 Deletion involving enhancer sequences within the gene DYNC1I1 presents with intellectual disability and split hand-split foot malformation with decreased penetrance Delgado, S. and Velinov, M., 2015 [75]
39 A large genomic deletion leads to enhancer adoption by the lamin B1 gene: a second path to autosomal dominant adult-onset demyelinating leukodystrophy (ADLD) Giorgio, E. et al., 2015 [76]
40 Multiple functional variants in long-range enhancer contribute to the risk of SNP rs965513 in thyroid cancer He, H. et al., 2015 [77]
41 Loss of TET2 in hematopoietic cells leads to DNA hypermethylation of active enhancers and induction of leukemogenesis Rasmussen et al., 2015 [78]
42 The Transcriptional cofactor TRIM33 prevents apoptosis B lymphoblastic leukemia by deactivating a single enhancer Wang et al., 2015 [79]
43 Super-enhancers delineate disease-associated regulatory nodes in T cells Vahedi et al., 2015 [44]
44 Identification of focally amplified lineage-specific super-enhancers in human epithelial cancer Zhang, X. et al., 2016 [80]
45 Role of non-coding sequence variants in cancer Khurana, E. et al., 2016 [81]
46 Ever-changing landscapes: transcriptional enhancers in development and evolution Long, H.K. et al., 2016 [82]
47 Genetic Predisposition to Chronic Lymphocytic Leukemia is mediated by a BMF Super-Enhancer Polymorphisms Kandaswamy et al., 2016 [83]
48 DNMT3A Loss drives Enhancer Hypomethylation in FLT3-ITD-Associated Leukemias Yang et al., 2016 [84]
49 Epigenomic profiling of primary gastric adenocarcinoma reveals super-enhancer heterogeneity Ooi et al., 2016 [43]
50 Parkinson associated risk variants in distal enhancers of a-syncuclein modulates target gene expression Soldner et al., 2016 [41]
51 Hotspots of aberrant enhancer activity punctuate the colorectal cancer epigenome Cohen, A.J. et al., 2017 [30]
52 Composition and dosage of a multipartite enhancer cluster control developmental expression of Ihh (Indian hedgehog) Will, A.J. et al., 2017 [85]
53 Superenhancer Analysis Defines Novel Epigenomic Subtypes of Non-APL AML, including an RARaalpha Dependency Targetable by SY-1425, a Potent and Selective RARalpha Agonist MCKeown, M.R. et al., 2017 [86]
54 Enhancer profiling identifies critical cancer genes and characterize cell identity in adult T-cell leukemia Wong et al., 2017 [87]
55 APOBEC signature mutation generate an oncogenic enhancer that drives LMO1 expression in T-ALL Leukemia Li et al., 2017 [88]

References

  1. Banerji, J.; Rusconi, S.; Schaffner, W. Expression of a beta-globin gene is enhanced by remote SV40 DNA sequences. Cell 1981, 27, 299–308.
  2. Cui, K.; Zang, C.; Roh, T.-Y.; Schones, D.E.; Childs, R.W.; Peng, W.; Zhao, K. Chromatin signatures in multipotent human hematopoietic stem cells indicate the fate of bivalent genes during differentiation. Cell Stem Cell 2009, 4, 80–93.
  3. Geyer, P.K.; Green, M.M.; Corces, V.G. Tissue-specific transcriptional enhancers may act in trans on the gene located in the homologous chromosome: The molecular basis of transvection in Drosophila. EMBO J. 1990, 9, 2247–2256.
  4. Lomvardas, S.; Barnea, G.; Pisapia, D.J.; Mendelsohn, M.; Kirkland, J.; Axel, R. Interchromosomal interactions and olfactory receptor choice. Cell 2006, 126, 403–413.
  5. Kron, K.J.; Bailey, S.D.; Lupien, M. Enhancer alterations in cancer: A source for a cell identity crisis. Genome Med. 2014, 6, 77.
  6. Buecker, C.; Wysocka, J. Enhancers as information integration hubs in development: Lessons from genomics. Trends Genet. 2012, 28, 276–284.
  7. Natoli, G. Maintaining Cell Identity through Global Control of Genomic Organization. Immunity 2010, 33, 12–24.
  8. He, A.; Kong, S.W.; Ma, Q.; Pu, W.T. Co-occupancy by multiple cardiac transcription factors identifies transcriptional enhancers active in heart. Proc. Natl. Acad. Sci. USA 2011, 108, 5632–5637.
  9. Song, L.; Zhang, Z.; Grasfeder, L.L.; Boyle, A.P.; Giresi, P.G.; Lee, B.-K.; Sheffield, N.C.; Gräf, S.; Huss, M.; Keefe, D.; et al. Open chromatin defined by DNaseI and FAIRE identifies regulatory elements that shape cell-type identity. Genome Res. 2011, 21, 1757–1767.
  10. Ostuni, R.; Piccolo, V.; Barozzi, I.; Polletti, S.; Termanini, A.; Bonifacio, S.; Curina, A.; Prosperini, E.; Ghisletti, S.; Natoli, G. Latent Enhancers Activated by Stimulation in Differentiated Cells. Cell 2013, 152, 157–171.
  11. Spitz, F.; Furlong, E.E.M. Transcription factors: From enhancer binding to developmental control. Nat. Rev. Genet. 2012, 13, 613–626.
  12. Herz, H.-M. Enhancer deregulation in cancer and other diseases. Bioessays 2016, 38, 1003–1015.
  13. Creyghton, M.P.; Cheng, A.W.; Welstead, G.G.; Kooistra, T.; Carey, B.W.; Steine, E.J.; Hanna, J.; Lodato, M.A.; Frampton, G.M.; Sharp, P.A.; et al. Histone H3K27ac separates active from poised enhancers and predicts developmental state. Proc. Natl. Acad. Sci. USA 2010, 107, 21931–21936.
  14. Heinz, S.; Benner, C.; Spann, N.; Bertolino, E.; Lin, Y.; Laslo, P.; Cheng, J.X.; Murre, C.; Singh, H.; Glass, C.K. Simple combinations of lineage-determining transcription factors prime cis-regulatory elements required for macrophage and B cell identities. Mol. Cell 2010, 38, 576–589.
  15. Xu, J.; Shao, Z.; Glass, K.; Bauer, D.E.; Pinello, L.; Van Handel, B.; Hou, S.; Stamatoyannopoulos, J.A.; Mikkola, H.K.; Yuan, G.-C.; et al. Combinatorial assembly of developmental stage-specific enhancers controls gene expression programs during human erythropoiesis. Dev. Cell 2012, 23, 796–811.
  16. Choukrallah, M.-A.; Song, S.; Rolink, A.G.; Burger, L.; Matthias, P. Enhancer repertoires are reshaped independently of early priming and heterochromatin dynamics during B cell differentiation. Nat. Commun. 2015, 6, 8324.
  17. Loke, J.; Assi, S.A.; Imperato, M.R.; Ptasinska, A.; Cauchy, P.; Grabovska, Y.; Soria, N.M.; Raghavan, M.; Delwel, H.R.; Cockerill, P.; et al. RUNX1-ETO and RUNX1-EVI1 Differentially Reprogram the Chromatin Landscape in t(8;21) and t(3;21) AML. Cell Rep. 2017, 19, 1654–1668.
  18. Prange, K.; Mandoli, A.; Kuznetsova, T.; Wang, S.-Y.; Sotoca, A.M.; Marneth, A.E.; Van Der Reijden, B.A.; Stunnenberg, H.G.; Martens, J.H.A. MLL-AF9 and MLL-AF4 oncofusion proteins bind a distinct enhancer repertoire and target the RUNX1 program in 11q23 acute myeloid leukemia. Oncogene 2017, 36, 3346–3356.
  19. Gröschel, S.; Sanders, M.A.; Hoogenboezem, R.; de Wit, E.; Bouwman, B.; Erpelinck, C.; Van Der Velden, V.H.; Havermans, M.; Avellino, R.; Van Lom, K.; et al. A single oncogenic enhancer rearrangement causes concomitant EVI1 and GATA2 deregulation in leukemia. Cell 2014, 157, 369–381.
  20. Rada-Iglesias, A.; Bajpai, R.; Swigut, T.; Brugmann, S.A.; Flynn, R.A.; Wysocka, J. A unique chromatin signature uncovers early developmental enhancers in humans. Nature 2011, 470, 279–283.
  21. Zentner, G.E.; Tesar, P.J.; Scacheri, P.C. Epigenetic signatures distinguish multiple classes of enhancers with distinct cellular functions. Genome Res. 2011, 21, 1273–1283.
  22. Cruz-Molina, S.; Respuela, P.; Tebartz, C.; Kolovos, P.; Nikolić, M.; Fueyo, R.; van Ijcken, W.; Grosveld, F.; Frommolt, P.; Bazzi, H.; et al. PRC2 Facilitates the Regulatory Topology Required for Poised Enhancer Function during Pluripotent Stem Cell Differentiation. Cell Stem Cell 2017, 20, 689–705.e9.
  23. Karnuta, J.M.; Scacheri, P.C. Enhancers: Bridging the gap between gene control and human disease. Hum. Mol. Genet. 2018, 27, R219–R227.
  24. Whyte, W.A.; Orlando, D.A.; Hnisz, D.; Abraham, B.; Lin, C.Y.; Kagey, M.H.; Rahl, P.B.; Lee, T.I.; Young, R.A. Master Transcription Factors and Mediator Establish Super-Enhancers at Key Cell Identity Genes. Cell 2013, 153, 307–319.
  25. Hnisz, D.; Abraham, B.; Lee, T.I.; Lau, A.; Saint-André, V.; Sigova, A.A.; Hoke, H.A.; Young, R.A. Super-enhancers in the control of cell identity and disease. Cell 2013, 155, 934–947.
  26. Parker, S.C.J.; Stitzel, M.L.; Taylor, L.; Orozco, J.M.; Erdos, M.R.; Akiyama, J.A.; van Bueren, K.L.; Chines, P.S.; Narisu, N.; Black, B.; et al. Chromatin stretch enhancer states drive cell-specific gene regulation and harbor human disease risk variants. Proc. Natl. Acad. Sci. USA 2013, 110, 17921–17926.
  27. Corradin, O.; Saiakhova, A.; Akhtar-Zaidi, B.; Myeroff, L.; Willis, J.; Dot}Lari, R.C.-S.; Lupien, M.; Markowitz, S.; Scacheri, P.C. Combinatorial effects of multiple enhancer variants in linkage disequilibrium dictate levels of gene expression to confer susceptibility to common traits. Genome Res. 2014, 24, 1–13.
  28. Pasquali, L.; Gaulton, K.; Rodriguez-Segui, S.A.; Mularoni, L.; Miguel-Escalada, I.; Akerman, I.; Tena, J.J.; Morán, I.; Gómez-Marín, C.; van de Bunt, M.; et al. Pancreatic islet enhancer clusters enriched in type 2 diabetes risk-associated variants. Nat. Genet. 2014, 46, 136–143.
  29. Lovén, J.; Hoke, H.A.; Lin, C.Y.; Lau, A.; Orlando, D.A.; Vakoc, C.R.; Bradner, J.E.; Lee, T.I.; Young, R.A. Selective Inhibition of Tumor Oncogenes by Disruption of Super-Enhancers. Cell 2013, 153, 320–334.
  30. Cohen, A.J.; Saiakhova, A.; Corradin, O.; Luppino, J.; Lovrenert, K.; Bartels, C.F.; Morrow, J.J.; Mack, S.C.; Dhillon, G.; Beard, L.; et al. Hotspots of aberrant enhancer activity punctuate the colorectal cancer epigenome. Nat. Commun. 2017, 8, 14400.
  31. Sur, I.; Taipale, J. The role of enhancers in cancer. Nat. Rev. Cancer 2016, 16, 483–493.
  32. Heyn, H.; Vidal, E.; Ferreira, H.J.; Vizoso, M.; Sayols, S.; Gomez, A.; Moran, S.; Boque-Sastre, R.; Guil, S.; Martinez-Cardus, A.; et al. Epigenomic analysis detects aberrant super-enhancer DNA methylation in human cancer. Genome Biol. 2016, 17, 1–16.
  33. Akhtar-Zaidi, B.; Cowper-Sal, R.; Corradin, O.; Saiakhova, A.; Bartels, C.F.; Balasubramanian, D.; Myeroff, L.; Lutterbaugh, J.; Jarrar, A.; Kalady, M.F.; et al. Epigenomic enhancer profiling defines a signature of colon cancer. Science 2012, 336, 736–739.
  34. Aran, D.; Sabato, S.; Hellman, A. DNA methylation of distal regulatory sites characterizes dysregulation of cancer genes. Genome Biol. 2013, 14, R21.
  35. Schaub, M.A.; Boyle, A.P.; Kundaje, A.; Batzoglou, S.; Snyder, M. Linking disease associations with regulatory information in the human genome. Genome Res. 2012, 22, 1748–1759.
  36. Maurano, M.T.; Humbert, R.; Rynes, E.; Thurman, R.E.; Haugen, E.; Wang, H.; Reynolds, A.P.; Sandstrom, R.; Qu, H.; Brody, J.; et al. Systematic localization of common disease-associated variation in regulatory DNA. Science 2012, 337, 1190–1195.
  37. Albert, F.W.; Kruglyak, L. The role of regulatory variation in complex traits and disease. Nat. Rev. Genet. 2015, 16, 197–212.
  38. Nicolae, D.L.; Gamazon, E.; Zhang, W.; Duan, S.; Dolan, M.E.; Cox, N.J. Trait-associated SNPs are more likely to be eQTLs: Annotation to enhance discovery from GWAS. PLoS Genet. 2010, 6, e1000888.
  39. Arda, H.E.; Tsai, J.; Rosli, Y.R.; Giresi, P.; Bottino, R.; Greenleaf, W.J.; Chang, H.Y.; Kim, S.K. A Chromatin Basis for Cell Lineage and Disease Risk in the Human Pancreas. Cell Syst. 2018, 7, 310–322.e4.
  40. Pham, T.-H.; Benner, C.; Lichtinger, M.; Schwarzfischer, L.; Hu, Y.; Andreesen, R.; Chen, W.; Rehli, M. Dynamic epigenetic enhancer signatures reveal key transcription factors associated with monocytic differentiation states. Blood 2012, 119, e161–e171.
  41. Soldner, F.; Stelzer, Y.; Shivalila, C.S.; Abraham, B.; Latourelle, J.C.; Barrasa, M.I.; Goldmann, J.; Myers, R.H.; Young, R.A.; Jaenisch, R. Parkinson-associated risk variant in distal enhancer of α-synuclein modulates target gene expression. Nature 2016, 533, 95–99.
  42. Weedon, M.; International Pancreatic Agenesis Consortium; Cebola, I.D.S.; Patch, A.-M.; Flanagan, S.; De Franco, E.; Caswell, R.; Rodriguez-Segui, S.A.; Shaw-Smith, C.; Cho, C.H.-H.; et al. Recessive mutations in a distal PTF1A enhancer cause isolated pancreatic agenesis. Nat. Genet. 2014, 46, 61–64.
  43. Ooi, W.F.; Xing, M.; Xu, C.; Yao, X.; Ramlee, M.K.; Lim, M.C.; Cao, F.; Lim, K.; Babu, D.; Poon, L.-F.; et al. Epigenomic profiling of primary gastric adenocarcinoma reveals super-enhancer heterogeneity. Nat. Commun. 2016, 7, 12983.
  44. Vahedi, G.; Kanno, Y.; Furumoto, Y.; Jiang, K.; Parker, S.C.J.; Erdos, M.R.; Davis, S.; Roychoudhuri, R.; Restifo, N.P.; Gadina, M.; et al. Super-enhancers delineate disease-associated regulatory nodes in T cells. Nature 2015, 520, 558–562.
  45. Taub, R.; Kirsch, I.; Morton, C.; Lenoir, G.; Swan, D.; Tronick, S.; Aaronson, S.; Leder, P. Translocation of the c-myc gene into the immunoglobulin heavy chain locus in human Burkitt lymphoma and murine plasmacytoma cells. Proc. Natl. Acad. Sci. USA. 1982, 79, 7837–7841.
  46. Lettice, L.A.; Heaney, S.J.; Purdie, L.A.; Li, L.; De Beer, P.; Oostra, B.A.; Goode, D.; Elgar, G.; Hill, R.E.; De Graaff, E. A long-range Shh enhancer regulates expression in the developing limb and fin and is associated with preaxial polydactyly. Hum. Mol. Genet. 2003, 12, 1725–1735.
  47. Loots, G.G.; Kneissel, M.; Keller, H.; Baptist, M.; Chang, J.; Collette, N.; Ovcharenko, D.; Plajzer-Frick, I.; Rubin, E.M. Genomic deletion of a long-range bone enhancer misregulates sclerostin in Van Buchem disease. Genome Res. 2005, 15, 928–935.
  48. Emison, E.S.; McCallion, A.S.; Kashuk, C.S.; Bush, R.T.; Grice, E.; Lin, S.; Portnoy, M.E.; Cutler, D.J.; Green, E.D.; Chakravarti, A. A common sex-dependent mutation in a RET enhancer underlies Hirschsprung disease risk. Nature 2005, 434, 857–863.
  49. Rahimov, F.; Program, N.C.S.; Marazita, M.L.; Visel, A.; Cooper, M.E.; Hitchler, M.J.; Rubini, M.; Domann, F.; Govil, M.; Christensen, K.; et al. Disruption of an AP-2 alpha binding site in an IRF6 enhancer is associated with cleft lip. Nat. Genet. 2008, 40, 1341–1347.
  50. Jia, L.; Landan, G.; Pomerantz, M.; Jaschek, R.; Herman, P.; Reich, D.; Yan, C.; Khalid, O.; Kantoff, P.; Oh, W.; et al. Functional enhancers at the gene-poor 8q24 cancer-linked locus. PLoS Genet. 2009, 5, e1000597.
  51. Pomerantz, M.M.; Ahmadiyeh, N.; Jia, L.; Herman, P.; Verzi, M.P.; Doddapaneni, H.; Beckwith, C.A.; Chan, J.A.; Hills, A.; Davis, M.; et al. The 8q24 cancer risk variant rs6983267 shows long-range interaction with MYC in colorectal cancer. Nat. Genet. 2009, 41, 882–884.
  52. Tuupanen, S.; Turunen, M.; Lehtonen, R.J.; Hallikas, O.; Vanharanta, S.; Kivioja, T.; Bjorklund, M.; Wei, G.-H.; Yan, J.; Niittymäki, I.; et al. The common colorectal cancer predisposition SNP rs6983267 at chromosome 8q24 confers potential to enhanced Wnt signaling. Nat. Genet. 2009, 41, 885–890.
  53. Sotelo, J.; Esposito, D.; Duhagon, M.A.; Banfield, K.; Mehalko, J.; Liao, H.; Stephens, R.M.; Harris, T.J.R.; Munroe, D.J.; Wu, X. Long-range enhancers on 8q24 regulate c-Myc. Proc. Natl. Acad. Sci. USA. 2010, 107, 3001–3005.
  54. Wasserman, N.F.; Aneas, I.; Nobrega, M.A. An 8q24 gene desert variant associated with prostate cancer risk confers differential in vivo activity to a MYC enhancer. Genome Res. 2010, 20, 1191–1197.
  55. Lettice, L.A.; Daniels, S.; Sweeney, E.; Venkataraman, S.; Devenney, P.S.; Gautier, P.; Morrison, H.; Fantes, J.; Hill, R.E.; Fitzpatrick, D.R. Enhancer-adoption as a mechanism of human developmental disease. Hum. Mutat. 2011, 32, 1492–1499.
  56. Sur, I.K.; Hallikas, O.; Vähärautio, A.; Yan, J.; Turunen, M.; Enge, M.; Taipale, M.; Karhu, A.; Aaltonen, L.A.; Taipale, J. Mice Lacking a Myc Enhancer That Includes Human SNP rs6983267 Are Resistant to Intestinal Tumors. Science 2012, 338, 1360–1363.
  57. Laurell, T.; VanderMeer, J.E.; Wenger, A.M.; Grigelioniene, G.; Nordenskjöld, A.; Arner, M.; Ekblom, A.G.; Bejerano, G.; Ahituv, N.; Nordgren, A. A novel 13 base pair insertion in the sonic hedgehog ZRS limb enhancer (ZRS/LMBR1) causes preaxial polydactyly with triphalangeal thumb. Hum. Mutat. 2012, 33, 1063–1066.
  58. Smemo, S.; Campos, L.C.; Moskowitz, I.P.; Krieger, J.E.; Pereira, A.C.; Nobrega, M.A. Regulatory variation in a TBX5 enhancer leads to isolated congenital heart disease. Hum. Mol. Genet. 2012, 21, 3255–3263.
  59. Aran, D.; Hellman, A. DNA methylation of transcriptional enhancers and cancer predisposition. Cell 2013, 154, 11–13.
  60. Chapuy, B.; McKeown, M.R.; Lin, C.Y.; Monti, S.; Roemer, M.G.; Qi, J.; Rahl, P.B.; Sun, H.H.; Yeda, K.T.; Doench, J.G.; et al. Discovery and characterization of super-enhancer-associated dependencies in diffuse large B cell lymphoma. Cancer Cell 2013, 24, 777–790.
  61. Shi, J.; Whyte, W.A.; Zepeda-Mendoza, C.J.; Milazzo, J.P.; Shen, C.; Roe, J.-S.; Minder, J.L.; Mercan, F.; Wang, E.; Eckersley-Maslin, M.; et al. Role of SWI/SNF in acute leukemia maintenance and enhancer-mediated Myc regulation. Genes Dev. 2013, 27, 2648–2662.
  62. Bauer, D.E.; Kamran, S.; Lessard, S.; Xu, J.; Fujiwara, Y.; Lin, C.; Shao, Z.; Canver, M.; Smith, E.C.; Pinello, L.; et al. An erythroid enhancer of BCL11A subject to genetic variation determines fetal hemoglobin level. Science 2013, 342, 253–257.
  63. Bhatia, S.; Bengani, H.; Fish, M.; Brown, A.; Divizia, M.T.; De Marco, R.; Damante, G.; Grainger, R.; Van Heyningen, V.; Kleinjan, D.A. Disruption of autoregulatory feedback by a mutation in a remote, ultraconserved PAX6 enhancer causes aniridia. Am. J. Hum. Genet. 2013, 93, 1126–1134.
  64. Weinhold, N.; Jacobsen, A.; Schultz, N.; Sander, C.; Lee, W. Genome-wide analysis of noncoding regulatory mutations in cancer. Nat. Genet. 2014, 46, 1160–1165.
  65. Herranz, D.; Ambesi-Impiombato, A.; Palomero, T.; Schnell, S.A.; Belver, L.; Wendorff, A.A.; Xu, L.; Castillo-Martin, M.; Llobet-Navás, D.; Cordon-Cardo, C.; et al. A NOTCH1-driven MYC enhancer promotes T cell development, transformation and acute lymphoblastic leukemia. Nat. Med. 2014, 20, 1130–1137.
  66. Northcott, P.A.; Lee, C.; Zichner, T.; Stütz, A.M.; Erkek, S.; Kawauchi, D.; Shih, D.J.H.; Hovestadt, V.; Zapatka, M.; Sturm, D.; et al. Enhancer hijacking activates GFI1 family oncogenes in medulloblastoma. Nature 2014, 511, 428–434.
  67. Lohan, S.; Spielmann, M.; Doelken, S.; Flottmann, R.; Muhammad, F.; Baig, S.; Wajid, M.; Hülsemann, W.; Habenicht, R.; Kjaer, K.; et al. Microduplications encompassing the Sonic hedgehog limb enhancer ZRS are associated with Haas-type polysyndactyly and Laurin-Sandrow syndrome. Clin. Genet. 2014, 86, 318–325.
  68. Seumois, G.; Chavez, L.; Gerasimova, A.; Lienhard, M.; Omran, N.; Kalinke, L.; Vedanayagam, M.; Ganesan, A.P.V.; Chawla, A.; Djukanović, R.; et al. Epigenomic analysis of primary human T cells reveals enhancers associated with T H 2 memory cell differentiation and asthma susceptibility. Nat. Immunol. 2014, 15, 777–788.
  69. Mansour, M.; Abraham, B.; Anders, L.; Berezovskaya, A.; Gutierrez, A.; Durbin, A.; Etchin, J.; Lawton, L.; Sallan, S.E.; Silverman, L.B.; et al. An oncogenic super-enhancer formed through somatic mutation of a noncoding intergenic element. Science 2014, 346, 1373–1377.
  70. Zhou, H.Y.; Katsman, Y.; Dhaliwal, N.K.; Davidson, S.; Macpherson, N.N.; Sakthidevi, M.; Collura, F.; Mitchell, J.A. A Sox2 distal enhancer cluster regulates embryonic stem cell differentiation potential. Genes Dev. 2014, 28, 2699–2711.
  71. Fortini, B.K.; Tring, S.; Plummer, S.J.; Edlund, C.K.; Moreno, V.; Bresalier, R.; Barry, E.L.; Church, T.R.; Figueiredo, J.C.; Casey, G. Multiple Functional Risk Variants in a SMAD7 Enhancer Implicate a Colorectal Cancer Risk Haplotype. PLoS ONE 2014, 9, e111914.
  72. Yamazaki, H.; Suzuki, M.; Otsuki, A.; Shimizu, R.; Bresnick, E.H.; Engel, J.D.; Yamamoto, M. A remote GATA2 hematopoietic enhancer drives leukemogenesis in inv(3)(q21;q26) by activating EVI1 expression. Cancer Cell 2014, 25, 415–427.
  73. Yashiro-Ohtani, Y.; Wang, H.; Zang, C.; Arnett, K.L.; Bailis, W.; Ho, Y.; Knoechel, B.; Lanauze, C.; Louis, L.; Forsyth, K.S.; et al. Long-range enhancer activity determines Myc sensitivity to Notch inhibitors in T cell leukemia. Proc. Natl. Acad. Sci. USA 2014, 111, E4946–E4953.
  74. Oldridge, D.A.; Wood, A.C.; Weichert-Leahey, N.; Crimmins, I.; Sussman, R.; Winter, C.; McDaniel, L.D.; Diamond, M.; Hart, L.S.; Zhu, S.; et al. Genetic predisposition to neuroblastoma mediated by a LMO1 super-enhancer polymorphism. Nature 2015, 528, 418–421.
  75. Delgado, S.; Velinov, M. 7q21.3 Deletion involving enhancer sequences within the gene DYNC1I1 presents with intellectual disability and split hand-split foot malformation with decreased penetrance. Mol. Cytogenet 2015, 8, 37.
  76. Giorgio, E.; Robyr, D.; Spielmann, M.; Ferrero, E.; Di Gregorio, E.; Imperiale, D.; Vaula, G.; Stamoulis, G.; Santoni, F.; Atzori, C.; et al. A large genomic deletion leads to enhancer adoption by the lamin B1 gene: A second path to autosomal dominant adult-onset demyelinating leukodystrophy (ADLD). Hum. Mol. Genet. 2015, 24, 3143–3154.
  77. He, H.; Li, W.; Liyanarachchi, S.; Srinivas, M.; Wang, Y.; Akagi, K.; Wang, Y.; Wu, D.; Wang, Q.; Jin, V.; et al. Multiple functional variants in long-range enhancer elements contribute to the risk of SNP rs965513 in thyroid cancer. Proc. Natl. Acad. Sci. USA 2015, 112, 6128–6133.
  78. Rasmussen, K.D.; Jia, G.; Johansen, J.V.; Pedersen, M.T.; Rapin, N.; Bagger, F.O.; Porse, B.T.; Bernard, O.A.; Christensen, J.; Helin, K. Loss of TET2 in hematopoietic cells leads to DNA hypermethylation of active enhancers and induction of leukemogenesis. Genes Dev. 2015, 29, 910–922.
  79. Wang, E.; Kawaoka, S.; Roe, J.-S.; Shi, J.; Hohmann, A.F.; Xu, Y.; Bhagwat, A.S.; Suzuki, Y.; Kinney, J.B.; Vakoc, C.R. The transcriptional cofactor TRIM33 prevents apoptosis in B lymphoblastic leukemia by deactivating a single enhancer. eLife 2015, 4, e06377.
  80. Zhang, X.; Choi, P.; Francis, J.M.; Imielinski, M.; Watanabe, H.; Cherniack, A.D.; Meyerson, M. Identification of focally amplified lineage-specific super-enhancers in human epithelial cancers. Nat. Genet. 2016, 48, 176–182.
  81. Khurana, E.; Fu, Y.; Chakravarty, D.; Demichelis, F.; Rubin, M.A.; Gerstein, M. Role of non-coding sequence variants in cancer. Nat. Rev. Genet. 2016, 17, 93–108.
  82. Long, H.K.; Prescott, S.L.; Wysocka, J. Ever-Changing Landscapes: Transcriptional Enhancers in Development and Evolution. Cell 2016, 167, 1170–1187.
  83. Kandaswamy, R.; Sava, G.; Speedy, H.E.; Beà, S.; Martín-Subero, J.I.; Studd, J.B.; Migliorini, G.; Law, P.; Suarez-Puente, X.; Garcia, D.M.; et al. Genetic Predisposition to Chronic Lymphocytic Leukemia Is Mediated by a BMF Super-Enhancer Polymorphism. Cell Rep. 2016, 16, 2061–2067.
  84. Yang, L.; Rodriguez, B.; Mayle, A.; Park, H.J.; Lin, X.; Luo, M.; Jeong, M.; Curry, C.V.; Kim, S.-B.; Ruau, D.; et al. DNMT3A Loss Drives Enhancer Hypomethylation in FLT3-ITD-Associated Leukemias. Cancer Cell 2016, 29, 922–934.
  85. Will, A.J.; Cova, G.; Osterwalder, M.; Chan, W.-L.; Wittler, L.; Brieske, N.; Heinrich, V.; de Villartay, J.; Vingron, M.; Klopocki, E.; et al. Composition and dosage of a multipartite enhancer cluster control developmental expression of Ihh (Indian hedgehog). Nat. Genet. 2017, 49, 1539–1545.
  86. McKeown, M.R.; Corces, M.R.; Eaton, M.L.; Fiore, C.; Lee, E.; Lopez, J.T.; Chen, M.W.; Smith, D.; Chan, S.M.; Koenig, J.L.; et al. Superenhancer Analysis Defines Novel Epigenomic Subtypes of Non-APL AML, Including an RARα Dependency Targetable by SY-1425, a Potent and Selective RARα Agonist. Cancer Discov. 2017, 7, 1136–1153.
  87. Wong, R.W.J.; Ngoc, P.C.T.; Leong, W.Z.; Yam, A.W.Y.; Zhang, T.; Asamitsu, K.; Iida, S.; Okamoto, T.; Ueda, R.; Gray, N.S.; et al. Enhancer profiling identifies critical cancer genes and characterizes cell identity in adult T-cell leukemia. Blood 2017, 130, 2326–2338.
  88. Li, Z.; Abraham, B.; Berezovskaya, A.; Farah, N.; Liu, Y.; Leon, T.E.; Fielding, A.; Tan, S.H.; Sanda, T.; Weintraub, A.S.; et al. APOBEC signature mutation generates an oncogenic enhancer that drives LMO1 expression in T-ALL. Leukemia 2017, 31, 2057–2064.
More
Information
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 376
Revisions: 3 times (View History)
Update Date: 12 Oct 2021
1000/1000