Open Access

Chromatin looping links gene expression to the assembly of transcription factories (Review)

  • Authors:
    • Bruno Perillo
    • Antimo Migliaccio
    • Gabriella Castoria
  • View Affiliations

  • Published online on: April 9, 2024     https://doi.org/10.3892/mmr.2024.13221
  • Article Number: 97
  • Copyright: © Perillo et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Genes are not randomly dispersed within the nuclear space, instead they occupy precise sites either with respect to the nuclear lamina as well as to each other. This observation stands at the basis of the today well accepted concept of nuclear territories where any chromosome shows reproducible spatial connections with a selection of others in a general picture that meets a functional criterion where genes that answer the same stimuli are grouped in the same sites. In fact, transcription is not visible widely dispersed throughout the nucleus but is gathered in several ‘granules’, called transcription factories that accommodates ~10 genes concurrently transcribed. This dynamic behavior of chromosomes is allowed by changes in chromatin plasticity that are governed by several classes of proteins that either modify its building or induce post‑translational modifications in the protein component of nucleosomes, triggering formation of chromosome loops that modify the location of specific sites along the DNA strand. For example, transcription associated to nuclear receptors benefits of the generation of nuclear ROS that induce nicks following activation of the DNA repair apparatus that enhance helix unfolding and chromosome bridging. In the present review, the role that protocols facing elucidation of chromosome architecture are playing and will play in the near future were highlighted in order to investigate composition of the transcription factories assembled in response of a specific trigger: The estrogen‑sensitive transcription was cited but the authors are convinced that the same portrait will be observed with a multitude of (if not all) other stimuli.

Introduction

The nucleus of human cells harbors 46 chromosomes that are organized into smaller domains, which enable packaging of different structural subunits devoted to different functions. Chromosomes occupy specific locations within the nuclear space, termed chromosome territories, which are further subdivided into chromosomal compartments [topologically associated domains (TADs)] that exhibit specific associations between promoter and enhancer sites (1). It is now universally accepted that location of co-regulated genes in different chromosomes is not casual, as they occupy the same regions within the nuclear space in different cells (2): This painting lies at the base of the notion of ‘chromosome territories’ where the spatial correlation between chromosomes changes throughout cell life (3). In this model, any gene exhibits different neighbors depending on the specific phase of the cell cycle and, in particular, on its activation state, indicating the presence of a specific molecular apparatus that drives chromatin mobility (4). As a corollary of chromosome territories, it has been also demonstrated that gene expression is grouped in several sites, called ‘transcription factories’, where genes that share the activating stimuli are concurrently transcribed (Fig. 1) (5,6). Physically, transcription factories can be viewed as nuclear granules with a diameter of roughly 50–100 nm reaching a number that varies between the different cell types and ranges from few hundreds to ~30,000 (7,8). They behave as chromatin hubs where at least two different active RNA polymerases synthesize RNA on two different targets (9,10), and harbor multi-protein complexes that reach local concentrations sufficient to complete ordered expression of co-regulated genes (11). In particular, it has been calculated that any factory contains 8–10 active genes on average (12,13).

Chromatin looping and distribution of genes in different territories and transcription factories

The molecular mechanism supporting co-localization of genes within the same factory needs to be investigated in deeper detail; however, it is essentially based on the achievement of a specific flexibility by chromatin that allows creation of several loops able to change location of genes with respect to each other along a chromatin fiber. On this regard, it could be hypothesized that large-scale random movements of chromatin may allow co-activated genes to meet the factories, triggering the multistep process that drives their ordered recruitment to the factory (14,15). This process implies the involvement of different classes of proteins such as chromatin modifiers, transcription factors and, notably, looping factors. Gene looping calls for particular consideration as it governs the precise apposition of enhancers and target genes mainly through different mechanisms where architectural DNA binding proteins contribute to looping by interacting with general or cell-specific factors (16,17). Moreover, occasionally the process involves the direct activity of general transcription factors including the mediator complex and cohesin, that help the looping between enhancers and promoters (18), whereas some other transcription factors, such as ZNF143 and YY1 have been evidenced to enhance chromatin loop formation in order to drive localization of specific loci within the nucleus (19,20). In addition, enhancer and non-coding RNA molecules have been deeply involved in looping between enhancer and promoter sites of specific genes, bringing these two elements into physical proximity in order to allow correct transcriptional processing and indicating chromatin looping as one of the features subjected to enhancer RNA control (2123). Finally, gene looping also concerns in most cases the 3′-end polyadenylation sites of genes that bridge with promoters and has been revealed to be essential for a correct transcriptional output (24). Interestingly, looping factors make a direct contribution to the recruitment of a specific gene to the factory through the establishment of protein bridges with the transcription factors already located in shared factories (25), governing, in this way, the unequal distribution of transcription throughout the nucleus in order to gain optimized gene expression (Fig. 1) (26).

The separation of nuclear space in chromosome territories is not inelastic and changes in response to environmental stimuli that shape compartmentalization of chromatin in parallel with restriction of the developmental progression (27). In particular, the genome appears widely dispersed within the nucleus in pluripotent stem cells, probably due to the hyper-dynamic nature of looping of chromatin fibers present in these cells (28). However, during lineage specification, a dynamic equilibrium between chromatin mobility and the transcriptional noise underlying specific gene silencing must be reached (29,30). As the specialization process progresses, chromatin compaction increases with large zones of the nucleus devoid of DNA that accumulates along the nuclear envelope and near the nucleolus: An image that correlates with the majority of genes transcriptionally repressed and consequent appearance of several Lamina-Associated-Domains where inhibitory markers prevail (31). These states of chromatin folding are acquired during cell type specialization and are usually reversible depending upon the environmental signal: In this regard an original example that underlines the role played by a specific cell function, is represented by the rod photoreceptor cells where compaction of chromosomes can be observed at the center of the nuclear space instead of periphery. In fact, in these cells, chromatin functions as a physical barrier to the scattering of light when passing through, and its nuclear distribution appears to be in accordance with the regulation of gene expression (32). The development of the concept of Transcription Factory and the characterization of some of such factories shed light on how the transcription factors drive the topological genome reorganization in the context of the control of gene expression. The authors' experience has been focused essentially on steroid receptors, in particular estrogen and retinoic acid receptors, but other examples of TFs mediating the assembly of transcription factories include Nanog (33) and Sox2 (34) in differentiation of pluripotent stem cells, Klf1 (35) in erythrocytes differentiation, Pax3 (36) and MyoD (37) during myogenesis, Pax5 (38) in B cell differentiation.

It is also important to clarify that the transcription factories represent by themselves well-defined entities, but more insight into three-dimensional chromatin changes associated with the activation of transcription could arise from the deeper knowledge about nuclear structures such as the membrane-less organelle (MLOs) (39). Evidence exists suggesting that nuclear condensate formation could be involved in the regulation of various aspects of gene expression, as numerous transcription factors, such as the steroid receptors, undergo Liquid-Liquid Phase Separation, the process leading to MLOs formation. Androgen receptor (AR), estrogen receptor and glucocorticoid receptor (GR) condensates were observed in the presence of the Mediator Complex subunit 1. The formation of condensates is subordinated to the interaction of the receptor with specific chromatin regions in the nucleus. The ligand-bound steroid receptors are translocated to the nucleus, where they form transcriptionally active foci. It has been reported that AR and GR foci are endowed with properties of MLOs (40). Thus, it is conceivable that formation of MLOs including steroid receptor-chromatin complexes contributes to stabilization of the transcription factories architecture and an integrated analysis of both processes could provide comprehensive insight into the regulation of gene expression.

On one extreme side of cellular differentiation and growth stands cancer, where altered nuclear architecture shows an irregular occupancy of the nucleus by chromosomes with altered compaction of chromatin and its DNA content. In particular, cancer cells display changes in the appearance and number of the nucleoli, the first described transcription factories assembled in proximity of the nucleolar organizing regions (41). In fact, an increase in the number of nucleoli parallels a decrease in the stability of contacts between chromosomes carrying rRNA loci, and consequent changes in nuclear organization that may impact adaptive responses. In this regard, cells have evolved a complex machinery to preserve the organization of nuclear space in which chromatin mobility is balanced by factors that restrain its amplitude; however, in transformed cells alterations in bridging between regulatory elements such as enhancers or insulators with promoters are well documented (42). An increase in mobility of cancer genome may also be presumably responsible for the increased gene expression with consequent phenotypic variability among cancer cells, functioning as the base of tumor progression due to altered contacts with the transcription factors that normally govern gene expression (43). As an example, translocations of Myc and Igh genes, most frequently found in plasmacytoma cells (44), share the same factories presumably because they require similar transcription factors that start the process of translocation (45,46). The observed increase in genome mobility may be due, at least in some cases, to changes in the activity of genome organizer proteins such as SATB1 that has been involved in the emergence of aggressive tumor phenotypes (47). Finally, a spatial relation between the three-dimensional architecture of the genome and chromosomal alterations has been reported in cancer cells where proximity of regions transcribed at the same time may be used to predict variations in gene-copy number observed frequently in malignant cells (48,49).

Experimental strategies to highlight transcriptional looping

Chromosome conformation capture (3C)

It is well established that transcription is governed by post-translational modifications at the N-terminal tails of the histone octamer which represents the major protein component of nucleosomes, the fundamental subunit of chromatin. Such modifications follow a precise code (50), and are especially aimed at governing the interaction of transcription factors with the DNA located at the promoter regions of genes to be activated and/or repressed (51,52). Among these modifications, methylation of lysine 4 (K4) and demethylation of lysine 9 (K9) in histone H3, the most protruding tail from the octamer histone disc, are mostly involved in gene activation (53,54). In particular, demethylation of H3K9 is catalyzed by specific demethylases recruited to the regulatory sites of activated genes (55), and produces nuclear reactive oxygen species (ROS) that induce single-strand DNA breaks that rule chromatin plasticity, easing the productive transcriptional output (56,57).

In order to analyze the involvement of gene folding in the expression of responsive genes induced by estrogens, the previously described strategy ‘Capturing chromosome conformation’, also known as ‘3C’ was first followed (58). In synthesis, the experimental design is based on the quantification of contact frequencies at any time between two loci distant on linear DNA, as revealed by quantitative polymerase chain reaction (qPCR) amplifications yielding incomparably richer data with respect to previous strategies based on light microscopy, and providing deeper insight into the spatial organization of specific loci (Fig. 2).

The main steps of 3C experimental design may be summarized as follows: i) cross-linking of chromatin using a soluble fixative agent such as formaldehyde to generate covalent bonds between the sites on DNA bridged by proteins; ii) isolation and digestion of chromatin with selected restriction enzymes; iii) ligation of sticky ends of digested fragments at low DNA concentration to favor intramolecular over intermolecular ligations; iv) reverse of cross-linking to obtain purified DNA in order to interrogate the rearranged fragments by qPCR using locus-specific primers encompassing fragments of the sites supposed to be bridged under the investigated experimental conditions; and v) comparison of the PCR results from cross-linked templates with control templates without bridging, to discriminate non-specific ligations. In principle, the cross-linking frequency between two sites on DNA is inversely proportional to their relative distance unless they are put in spatial contact by a particular protein factor. Therefore, a picture of the three-dimensional (3D) architecture of a particular locus may be deduced by comparing linking frequencies from cross-linked chromatin to the control (Fig. 2).

Historically, yeast chromosome III has been the first to be analyzed, and its 3D conformation has revealed proximity between the telomeres that conferred the chromosome the form of a sort of ring (58). Next, 3C experimental approach was adapted for analysis of mammary cells and, even though it can be especially used to detect contacts between sites spanning only few hundred kilobases (59), it confirmed the existence of dynamic chromatin loops between promoters/enhancers and their target genes, conferring chromatin a peculiar conformation dependent on the transcriptional status (6062).

Upon 3C, it was demonstrated that the estrogen responsive element located ~1.5 kb downstream from the transcription start site of the estrogen responsive gene bcl-2 bridged through a complex containing the estrogen receptor with the transcription complex located onto the promoter of the same gene. Moreover, looping between the promoter at the 5′-end, and the 3′-end where the polyadenylation site is located, has been widely assessed, supporting the concept that 3′ end-processing factors bridge with the transcriptional machinery (63,64).

DNA-picked chromatin (DPC)

3C certainly represents a classical, strong tool to identify gene looping; nevertheless, it requires the existence of suitable restriction sites strategically located along the genes under investigation and setting up the right balance between the concentration of DNA and the restriction enzyme to enhance intramolecular ligations. To try to overcome these bottlenecks, a novel experimental strategy conceived by introducing several changes into the method designed to perform proteomic analysis of multiprotein complexes assembled on chromatin (65) was proposed. This strategy, called DPC (DNA-picked chromatin) by the authors, consisted in cross-linking chromatin from cells challenged or not with estrogens, DNA cleavage to 500–600 base-pair fragments by sonication, and hybridization with a biotinylated oligonucleotide probe complementary to the site under investigation, followed by capture on magnetic streptavidin beads (Fig. 3). Presence of DNA loci bordering the site hybridized by the probe was, then, revealed by qPCR amplifications of the DNA extracted and purified from the eluted fragments after de-crosslinking. As a control, PCR reactions were carried out using chromatin before hybridization or naked DNA as template (input), and recovery of the probing region in retrieved DNA from each experimental condition was assumed as the baseline to be compared with the relative amount of retrieved associated sites in the same experimental point. Moreover, to solve troubles emerged by possible changes of spatial chromatin organization throughout transcriptional stimulation, the use of multiple probes was recommended. In summary, the strategy assumes that two sites can be co-captured by the probe only if they are bridged by proteins recruited upon the transcriptional trigger, independently from their distance on linear DNA, and can be imagined as a sort of chromatin immunoprecipitation (ChIP) where the antibody is substituted with the DNA probe (Fig. 3) (66).

Owing to DPC strategy, the spatial relationship between the enhancer and polyadenylation sites of bcl-2 gene upon estrogen addition was first analyzed and it was observed that the last was increased in the DNA purified after hybridization with the enhancer probe (65). More important, this technique allowed to assess the connection within the nuclear space between this gene, located on chromosome 18, and another estrogen-sensitive gene, RIZ, located on chromosome 1, and the authors were able to detect an increase of bcl-2 enhancer and polyadenylation sites when the rescued DNA hybridized with RIZ promoter was used as bait (65). Noteworthy, two different chromosomes, 1 and 18, have been demonstrated to occupy contiguous territories within the nuclear space and previous data by the authors suggested that this is dependent on assembly of the cognate transcription complex after hormone stimulation (67). It was also assessed that the demethylation of H3K9me2 is essential to the formation of transcription-induced looping of hormone-dependent genes (66).

Circular chromosome conformation capture (4C)

Having demonstrated the intimate correlation between changes induced on chromatin flexibility by transcription (through induction of single-strand breaks by an increase of ROS production due to the demethylase activity) and establishment of loops that allow allocation of concurrently transcribed genes within the same transcription factories, the further step is represented by the identification of the genes involved in the assembly of such factories, starting by a gene whose responsiveness to the investigated trigger is universally recognized (in the present case, the bait is represented by an estrogen-sensitive gene). In fact, as has been graphically represented in Fig. 1, if the estrogen responsive factory is imagined as a flower with the transcribed genes being the pistils and the intervening DNA the petals, then the goal should be the elucidation of the genes that form the factory/flower establishing contacts to the bait (Fig. 3).

To this aim, since either 3C as well as DPC focus on interactions between two already known loci (for this reason called ‘one versus one’) and can, then, be used exclusively to detect spatial interactions between already known regions (68), an-omics approach is needed, using the products generated by the 3C or DPC procedures as a library that needs further treatment. On this regard, a growing family of related strategies has been reported, among which the 3C-on-chip (69), and the open-ended 3C (70), mostly based on circularization of the generated chromatin fragments using a second restriction enzyme. Thus, the authors' attention will be now pointed to the 4C experimental approach that represents a paramount protocol to assess the involvement of co-responsive genes within the same transcription factory (a ‘one versus all’ approach) (71). In synthesis, the procedure may be summarized as follows: Cross-linked chromatin is treated to generate circular DNA molecules from restricted hybrid fragments by use of a second restriction enzyme under high ligase concentration and prolonged incubation times. After reversal of cross-linking, nested PCR primers spanning the opposite ends of the DNA site chosen as probe are used to amplify any sequence fallen in strict proximity with the bait (also called the ‘viewpoint’) and then, ligated. The amplified ligation products are sequenced to assess all the spatial partners of the gene (or a specific locus of it) under investigation, providing the identity of all the pistils of the flower (Fig. 4) (72).

Conclusions and future perspectives

4C has been demonstrated to represent a useful device to highlight either short-range interactions as well as long-range spatial cross-talks between very distant sites (73); therefore, it appears as the strategy of choice to obtain a complete detection of genes that co-localize with the estrogen-responsive gene PS2 after hormone challenge (74). The same strategy will presumably allow assessment of components of any of the transcription factories established by any transcriptional trigger throughout cellular life-span, either in physiological as well as pathological conditions.

In conclusion, the scope of this brief review was to emphasize the importance of studying the three-dimensional structure of chromatin with the hope of providing a succinct summary, based on the authors' experience of the evolution of the principal methods used to deepen knowledge about this issue; the respective advantages and limits have been highlighted in Table I.

Table I.

Although the 4C appears undoubtedly as the most performing method to discover and analyze new transcription factories, it should be reiterated that 3C and the less widespread DPC techniques still remain very effective tools when a specific target is under investigation.

Table I.

Although the 4C appears undoubtedly as the most performing method to discover and analyze new transcription factories, it should be reiterated that 3C and the less widespread DPC techniques still remain very effective tools when a specific target is under investigation.

TechniqueAdvantagesLimits
3CEasy to use, high sensitivity, inexpensivePossible unavailability of restriction sites, undesired structure alteration due to cross-linking, informative only for two known sites
DPCIndependent on availability of restriction sites, free of possible experimental artifacts, very fast, high sensitivity, negligible costsInformative only for two known sites, limited by the interference of some three-dimensional chromatin structure
4CAllows a genome-wide analysis of the transcription factoriesPossible unavailability of restriction sites, rather expensive, requires preliminary ‘in silico’ analysis

[i] 4C, circular chromosome conformation capture; DPC, DNA-picked chromatin; 3C, chromosome conformation capture.

Acknowledgements

Not applicable.

Funding

The present study was supported by The Italian Ministry of Health-Project of finalized research RF-2019-12368937 ‘Targeting novel insulin/IGF-driven signaling networks in breast cancer’ (grant no. CUP D69J21001840001).

Availability of data and materials

Data sharing is not applicable to this article, as no data sets were generated or analyzed during the current study.

Authors' contributions

BP conceived, wrote and edited the manuscript. AM discussed and revised the manuscript and prepared figures. GC discussed and revised the manuscript. All authors read and approved the final manuscript. Data authentication is not applicable.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

References

1 

Fritz AJ, Sehgal N, Piss A, Xu J and Berezney R: Chromosome territories and the global regulation of the genome. Genes Chromosomes Cancer. 58:407–426. 2019. View Article : Google Scholar : PubMed/NCBI

2 

Rajapakse I, Perlman MD, Scalzo D, Kooperberg C, Groudine M and Kosak ST: The emergence of lineage-specific chromosomal topologies from coordinate gene regulation. Proc Natl Acad Sci USA. 106:6679–6684. 2009. View Article : Google Scholar : PubMed/NCBI

3 

Cremer T, Cremer M, Dietzel S, Müller S, Solovei I and Fakan S: Chromosome territories-a functional nuclear landscape. Curr Opin Cell Biol. 18:307–316. 2006. View Article : Google Scholar : PubMed/NCBI

4 

Chuang CH, Carpenter AE, Fuchsova B, Johnson T, de Lanerolle P and Belmont AS: Long-range directional movement of an interphase chromosome site. Curr Biol. 18:825–831. 2006. View Article : Google Scholar : PubMed/NCBI

5 

Xu M and Cook PR: Similar active genes cluster in specialized transcription factories. J Cell Biol. 181:615–623. 2008. View Article : Google Scholar : PubMed/NCBI

6 

Sutherland H and Bickmore WA: Transcription factories: Gene expression in unions? Nat Rev Genet. 10:457–466. 2009. View Article : Google Scholar : PubMed/NCBI

7 

Pombo A, Hollinshead M and Cook PR: Bridging the resolution gap: Imaging the same transcription factories in cryosections by light and electron microscopy. J Histochem Cytochem. 47:471–480. 1999. View Article : Google Scholar : PubMed/NCBI

8 

Faro-Trindade I and Cook PR: A conserved organization of transcription during embryonic stem cell differentiation and in cells with high C value. Mol Biol Cell. 17:2910–2920. 2006. View Article : Google Scholar : PubMed/NCBI

9 

de Laat W and Grosveld F: Spatial organization of gene expression: The active chromatin hub. Chromosome Res. 11:447–459. 2003. View Article : Google Scholar : PubMed/NCBI

10 

Papantonis A and Cook PR: Transcription factories: Genome organization and gene regulation. Chem Rev. 13:8683–8705. 2013. View Article : Google Scholar : PubMed/NCBI

11 

Osborne CS, Chakalova L, Brown KE, Carter D, Horton A, Debrand E, Goyenechea B, Mitchell JA, Lopes S, Reik W and Fraser P: Active genes dynamically colocalize to shared sites of ongoing transcription. Nat Genet. 36:1065–1071. 2004. View Article : Google Scholar : PubMed/NCBI

12 

Faro-Trindade I and Cook PR: Transcription factories: Structures conserved during differentiation and evolution. Biochem Soc Trans. 34:1133–1137. 2006. View Article : Google Scholar : PubMed/NCBI

13 

Larkin JD, Cook PR and Papantonis A: Dynamic reconfiguration of long human genes during one transcription cycle. Mol Cell Biol. 32:2738–2747. 2012. View Article : Google Scholar : PubMed/NCBI

14 

Mahy NL, Perry PE and Bickmore WA: Gene density and transcription influence the localization of chromatin outside of chromosome territories detectable by FISH. J Cell Biol. 159:753–763. 2002. View Article : Google Scholar : PubMed/NCBI

15 

Chambeyron S and Bickmore WA: Chromatin decondensation and nuclear reorganization of the HoxB locus upon induction of transcription. Genes Dev. 18:1119–1130. 2004. View Article : Google Scholar : PubMed/NCBI

16 

Krivega I and Dean A: Enhancer and promoter interactions-long distance calls. Curr Opin Genet Dev. 22:1–7. 2012. View Article : Google Scholar

17 

Stadhouders R, Thongjuea S, Andrieu-Soler C, Palstra RJ, Bryne JC, van den Heuvel A, Stevens M, de Boer E, Kockx C, van der Sloot A, et al: Dynamic long-range chromatin interactions control Myb proto-oncogene transcription during erythroid development. EMBO J. 31:986–999. 2012. View Article : Google Scholar : PubMed/NCBI

18 

Kagey MH, Newman JJ, Bilodeau S, Zhan Y, Orlando DA, van Berkum NL, Ebmeier CC, Goossens J, Rahl PB, Levine SS, et al: Mediator and cohesion connect gene expression and chromatin architecture. Nature. 467:430–435. 2010. View Article : Google Scholar : PubMed/NCBI

19 

Zhou Q, Yu M, Tirado-Magallanes M, Li B, Kong L, Guo M, Tan ZH, Lee S, Chai L, Numata A, et al: ZNF143 mediates CTCF-bound promoter-enhancer loops required for murine hematopoietic stem and progenitor cell function. Nat Commun. 12:432021. View Article : Google Scholar : PubMed/NCBI

20 

Weintraub AS, Li CH, Zamudio AV, Sigova AA, Hannett NM, Day DS, Abraham BJ, Cohen MA, Nabet B, Buckley DL, et al: YY1 is a structural regulator of enhancer-promoter loops. Cell. 171:1573–1588. 2017. View Article : Google Scholar : PubMed/NCBI

21 

Jiao W, Chen Y, Song H, Li D, Mei H, Yang F, Fang E, Wang X, Huang K, Zheng L and Tong Q: HPSE enhancer RNA promotes cancer progression through driving chromatin looping and regulating hnRNPU/p300/EGR1/HPSE axis. Oncogene. 36:2728–2745. 2018. View Article : Google Scholar : PubMed/NCBI

22 

Hsieh CL, Fei T, Chen Y, Li T, Gao Y, Wang X, Sun T, Sweeney CJ, Lee GS, Chen S, et al: Enhancer RNAs participate in androgen receptor-driven looping that selectively enhances gene activation. Proc Natl Acad Sci USA. 111:7319–7324. 2014. View Article : Google Scholar : PubMed/NCBI

23 

Tan SH, Leong WZ, Ngoc PCT, Tan TK, Bertulfo FC, Lim MC, An O, Li Z, Yeoh AEJ, Fullwood MJ, et al: The enhancer RNA ARIEL activates the oncogenic transcriptional program in T-cell acute lymphoblastic leukemia. Blood. 134:239–251. 2019. View Article : Google Scholar : PubMed/NCBI

24 

He X, Khan AU, Cheng H, Pappas DL Jr, Hampsey M and Moore CL: Functional interactions between the transcription and the mRNA 3′ end processing machineries mediated by Ssu72 and Sub1. Genes Dev. 17:1030–1042. 2003. View Article : Google Scholar : PubMed/NCBI

25 

Deng W, Lee J, Wang H, Miller J, Reik A, Gregory PD, Dean A and Blobel GA: Controlling long-range genomic interactions at a native locus by targeted tethering of a looping factor. Cell. 149:1233–1244. 2012. View Article : Google Scholar : PubMed/NCBI

26 

Chubb JR and Bickmore WA: Considering nuclear compartmentalization in the light of nuclear dynamics. Cell. 112:403–406. 2003. View Article : Google Scholar : PubMed/NCBI

27 

Chubb JR, Boyle S, Perry P and Bickmore WA: Chromatin motion is constrained by association with nuclear compartments in human cells. Curr Biol. 12:439–445. 2002. View Article : Google Scholar : PubMed/NCBI

28 

Meshorer E, Yellajoshula D, George E, Scambler PJ, Brown DT and Misteli T: Hyper-dynamic plasticity of chromatin proteins in pluripotent embryonic stem cells. Dev Cell. 10:105–116. 2006. View Article : Google Scholar : PubMed/NCBI

29 

Turner BM: Open chromatin and hypertranscription in embryonic stem cells. Cell Stem Cell. 2:408–410. 2008. View Article : Google Scholar : PubMed/NCBI

30 

Chang HH, Hemberg M, Barahona M, Ingber DE and Huang S: Transcriptome-wide noise controls lineage choice in mammalian progenitor cells. Nature. 453:544–547. 2008. View Article : Google Scholar : PubMed/NCBI

31 

Guelen L, Pagie L, Brasset E, Meuleman W, Faza MB, Talhout W, Eussen BH, de Klein A, Wesels L, de Laat W and van Steensel B: Domain organization of human chromosomes revealed by mapping of nuclear lamina interactions. Nature. 453:948–951. 2008. View Article : Google Scholar : PubMed/NCBI

32 

Solovei I, Kreysing M, Lanctôt C, Kösem S, Peichl L, Cremer T, Guck J and Joffe B: Nuclear architecture of rod photoreceptor cells adapts to vision in mammalian evolution. Cell. 137:356–368. 2009. View Article : Google Scholar : PubMed/NCBI

33 

Nitzsche A, Paszkowski-Rogacz M, Matarese F, Janssen-Megens EM, Hubner NC, Schulz H, de Vries I, Ding L, Huebner N, Mann M, et al: RAD21 cooperates with pluripotency transcription factors in the maintenance of embryonic stem cell identity. PLoS One. 6:e194702011. View Article : Google Scholar : PubMed/NCBI

34 

Bertolini JA, Favaro R, Zhu Y, Pagin M, Ngan CY, Wong CH, Tjong H, Vermunt MW, Martynoga B, Barone C, et al: Mapping the global chromatin connectivity network for Sox2 function in neural stem cell maintenance. Cell Stem Cell. 24:462–476. 2019. View Article : Google Scholar : PubMed/NCBI

35 

Drissen R, Palstra RJ, Gillemans N, Splinter E, Grosveld F, Philipsen S and de Laat W: The active spatial organization of the beta-globin locus requires the transcription factor EKLF. Genes Dev. 18:2485–2490. 2004. View Article : Google Scholar : PubMed/NCBI

36 

Magli A, Baik J, Pota P, Cordero CO, Kwak IY, Garry DJ, Love PE, Dynlacht BD and Perlingeiro RCR: Pax3 cooperates with Ldb1 to direct local chromosome architecture during myogenic lineage specification. Nat Commun. 10:23162019. View Article : Google Scholar : PubMed/NCBI

37 

Dall'Agnese A, Caputo L, Nicoletti C, di Iulio J, Schmitt A, Gatto S, Diao Y, Ye Z, Forcato M, Perera R, et al: Transcription factor-directed re-wiring of chromatin architecture for somatic cell nuclear reprogramming toward trans-differentiation. Mol Cell. 76:453–472. 2019. View Article : Google Scholar : PubMed/NCBI

38 

Johanson TM, Lun ATL, Coughlan HD, Tan T, Smyth GK, Nutt SL and Allan RS: Transcription-factor-mediated supervision of global genome architecture maintains B cell identity. Nat Immunol. 19:1257–1264. 2018. View Article : Google Scholar : PubMed/NCBI

39 

Sołtys K and Ozyhar A: Transcription regulators and membrane less organelles challenges to investigate them. Int J Mol Sci. 22:127582021. View Article : Google Scholar : PubMed/NCBI

40 

Shrinivas K, Sabari BR, Coffey EL, Klein IA, Boija A, Zamudio AV, Schuijers J, Hannett NM, Sharp PA, Young RA and Chakraborty AK: Enhancer features that drive formation of transcriptional condensates. Mol Cell. 75:549–561.e7. 2019. View Article : Google Scholar : PubMed/NCBI

41 

Pederson T: The nucleolus. Cold Spring Harbor Perspect Biol. 3:a0003682011. View Article : Google Scholar

42 

Gondor A and Ohlsson R: Chromosome crosstalk in three dimensions. Nature. 461:212–217. 2009. View Article : Google Scholar : PubMed/NCBI

43 

Pujadas E and Feinberg AP: Regulated noise in the epigenetic landscape of development and disease. Cell. 148:1123–1131. 2012. View Article : Google Scholar : PubMed/NCBI

44 

Osborne CS, Chakalova L, Mitchell JA, Horton A, Wood AL, Bolland DJ, Corcoran AE and Fraser P: Myc dynamically and preferentially relocates to a transcription factory occupied by Igh. PLoS Biol. 5:e1922007. View Article : Google Scholar : PubMed/NCBI

45 

Misteli T: Higher-order genome organization in human disease. Cold Spring Harb Perspect Biol. 2:a0007942010. View Article : Google Scholar : PubMed/NCBI

46 

Kohwi-Shigematsu T, Poterlowicz K, Ordinario E, Han HJ, Botchkarev V and Kohwi Y: Genome organizing function of SATB1 in tumor progression. Semin Cancer Biol. 23:72–79. 2013. View Article : Google Scholar : PubMed/NCBI

47 

Wijchers PJ and de Laat W: Genome organization influences partner selection for chromosomal rearrangements. Trends Genet. 27:63–71. 2011. View Article : Google Scholar : PubMed/NCBI

48 

De S and Michor F: DNA replication timing and long-range DNA interactions predict mutational landscapes of cancer genomes. Nat Biotechnol. 29:1103–1108. 2011. View Article : Google Scholar : PubMed/NCBI

49 

Fudenberg G, Getz G, Meyerson M and Mirny LA: High order chromatin architecture shapes the landscape of chromosomal alterations in cancer. Nat Biotechnol. 29:1109–1113. 2011. View Article : Google Scholar : PubMed/NCBI

50 

Strahl BD and Allis CD: The language of covalent histone modifications. Nature. 403:41–45. 2000. View Article : Google Scholar : PubMed/NCBI

51 

Berger SL: The complex language of chromatin regulation during transcription. Nature. 447:407–412. 2007. View Article : Google Scholar : PubMed/NCBI

52 

Kouzarides T: Chromatin modifications and their functions. Cell. 128:693–705. 2007. View Article : Google Scholar : PubMed/NCBI

53 

Peters AH, Kubicek S, Mechtler K, O'Sullivan RJ, Derijck AAH, Perez-Burgos L, Kohlmaier A, Opravil S, Tachibana M, Shinkai Y, et al: Partitioning and plasticity of repressive histone methylation states in mammalian chromatin. Mol Cell. 12:1577–1589. 2003. View Article : Google Scholar : PubMed/NCBI

54 

Rice JC, Briggs SD, Ueberheide B, Barber CM, Shabanowitz J, Hunt DF, Shinkai Y and Allis CD: Histone methyltransferases direct different degrees of methylation to define distinct chromatin domains. Mol Cell. 12:1591–1598. 2003. View Article : Google Scholar : PubMed/NCBI

55 

Closs PAC, Christensen J, Agger K and Helin K: Erasing the methyl mark: Histone demethylases at the center of cellular differentiation and disease. Genes Dev. 22:1115–1140. 2008. View Article : Google Scholar : PubMed/NCBI

56 

Perillo B, Ombra MN, Bertoni A, Cuozzo C, Sacchetti S, Sasso A, Chiariotti L, Malorni A, Abbondanza C and Avvedimento EV: DNA oxidation as triggered by H3K9me2 demethylation drives estrogen-induced gene expression. Science. 319:202–206. 2008. View Article : Google Scholar : PubMed/NCBI

57 

Papantonis A and Cook PR: Genome architecture and the role of transcription. Curr Opin Cell Biol. 22:271–276. 2010. View Article : Google Scholar : PubMed/NCBI

58 

Dekker J, Rippe K, Dekker M and Kleckner N: Capturing chromosome conformation. Science. 295:1306–1311. 2002. View Article : Google Scholar : PubMed/NCBI

59 

Simonis M, Kooren J and de Laat W: An evaluation of 3C-based methods to capture DNA interactions. Nat Methods. 4:895–901. 2007. View Article : Google Scholar : PubMed/NCBI

60 

Murrell A, Heeson S and Reik W: Interaction between differentially methylated regions partitions the imprinted genes Igf2 and H19 into parent-specific chromatin loops. Nat Genet. 36:889–893. 2004. View Article : Google Scholar : PubMed/NCBI

61 

Gheldof N, Smith EM, Tabuchi TM, Koch CM, Dunham I, Stamatoyannopoulos JA and Dekker J: Cell-type-specific long-range looping interactions identify distant regulatory elements of the CFTR gene. Nucleic Acids Res. 38:4325–4336. 2010. View Article : Google Scholar : PubMed/NCBI

62 

Dekker J, Marti-Renom MA and Mirny LA: Exploring the three-dimensional organization of genomes: Interpreting chromatin interaction data. Nat Rev Genet. 14:390–403. 2013. View Article : Google Scholar : PubMed/NCBI

63 

O'Sullivan JM, Tan-Wong SM, Morillon A, Lee B, Coles J, Mellor J and Proudfoot NJ: Gene loops juxtapose promoters and terminators in yeast. Nat Genet. 36:1014–1018. 2004. View Article : Google Scholar : PubMed/NCBI

64 

Bentley DL: Rules of engagement: Co-transcriptional recruitment of pre-mRNA processing factors. Curr Opin Cell Biol. 17:251–256. 2005. View Article : Google Scholar : PubMed/NCBI

65 

Déjardin J and Kingston RE: Purification of proteins associated with specific genomic loci. Cell. 136:175–186. 2009. View Article : Google Scholar : PubMed/NCBI

66 

Abbondanza C, De Rosa C, Ombra MN, Aceto F, Medici N, Altucci L, Moncharmont B, Puca GA, Porcellini A, Avvedimento EV and Perillo B: Highlighting chromosome loops in DNA-picked chromatin (DPC). Epigenetics. 6:979–986. 2011. View Article : Google Scholar : PubMed/NCBI

67 

Bolzer A, Kreth G, Solovei I, Koehler D, Sarakoglu K, Fauth C, Muller S, Eils R, Cremer C, Speicher MR and Cremer T: Three-dimensional maps of all chromosomes in human male fibroblast nuclei and prometaphase rosettes. PLoS Biol. 3:e1572005. View Article : Google Scholar : PubMed/NCBI

68 

Han J, Zhang Z and Wang K: 3C and 3C-based techniques: The powerful tools for spatial genome organization deciphering. Mol Cytogen. 9:11–21. 2018.

69 

Simonis M, Klous P, Splinter E, Moshkin Y, Willemsen R, de Wit E, van Steensel B and de Laat W: Nuclear organization of active and inactive chromatin domains uncovered by chromosome conformation capture-on-chip (4C). Nat Genet. 38:1348–1354. 2006. View Article : Google Scholar : PubMed/NCBI

70 

Wurtele H and Chartrand P: Genome-wide scanning of HoxB1-associated loci in mouse ES cells using an open-ended chromosome conformation capture methodology. Chromosome Res. 14:477–495. 2006. View Article : Google Scholar : PubMed/NCBI

71 

Zhao Z, Tavoosidana G, Sjölinder M, Göndör A, Mariano P, Wang S, Kanduri C, Lezcano M, Sandhu KS, Singh U, et al: Circular chromosome conformation capture (4C) uncovers extensive networks of epigenetically regulated intra- and interchromosomal interactions. Nat Genet. 38:1341–1347. 2006. View Article : Google Scholar : PubMed/NCBI

72 

van de Werken HJ, Landan G, Holwerda SJ, Hoichman M, Klous P, Chachik R, Splinter E, Valdes-Quezada C, Oz Y, Bouwman BAM, et al: Robust 4C-seq data analysis to screen for regulatory DNA interactions. Nat Methods. 9:969–972. 2012. View Article : Google Scholar : PubMed/NCBI

73 

Symmons O, Pan L, Remeseiro S, Aktas T, Klein F, Huber W and Spitz F: The Shh topological domain facilitates the action of remote enhancers reducing the effects of genomic distances. Dev Cell. 5:529–543. 2016. View Article : Google Scholar : PubMed/NCBI

74 

Tomasetto C, Rockel N, Mattei MG, Fujita R and Rio MC: The gene encoding the human spasmolytic protein (SML1/hSP) is in 21q physically linked to the homologous breast cancer marker gene BCEI/pS2. Genomics. 13:1328–1330. 1992. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

June-2024
Volume 29 Issue 6

Print ISSN: 1791-2997
Online ISSN:1791-3004

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Perillo B, Migliaccio A and Castoria G: Chromatin looping links gene expression to the assembly of transcription factories (Review). Mol Med Rep 29: 97, 2024
APA
Perillo, B., Migliaccio, A., & Castoria, G. (2024). Chromatin looping links gene expression to the assembly of transcription factories (Review). Molecular Medicine Reports, 29, 97. https://doi.org/10.3892/mmr.2024.13221
MLA
Perillo, B., Migliaccio, A., Castoria, G."Chromatin looping links gene expression to the assembly of transcription factories (Review)". Molecular Medicine Reports 29.6 (2024): 97.
Chicago
Perillo, B., Migliaccio, A., Castoria, G."Chromatin looping links gene expression to the assembly of transcription factories (Review)". Molecular Medicine Reports 29, no. 6 (2024): 97. https://doi.org/10.3892/mmr.2024.13221