Open access

Introductory Chapter: Gene Expression in Eukaryotic Cells

Written By

Fumiaki Uchiumi

Submitted: 15 January 2022 Published: 05 October 2022

DOI: 10.5772/intechopen.103152

From the Edited Volume

Gene Expression

Edited by Fumiaki Uchiumi

Chapter metrics overview

161 Chapter Downloads

View Full Metrics

1. Introduction

“Central Dogma” explains how information from genes to proteins flows. Genes should be transcribed into messenger ribonucleic acids (mRNAs) in nuclei, then they are processed and delivered to the cytoplasm where they are translated into polypeptides (proteins). We, molecular biologists, know that gene expression in mammalian cells is controlled at multiple stages. First, gene expression is epigenetically regulated by chromatin structures, which depend on deoxyribonucleic acid (DNA) methylation and histone modifications. Then, transcription initiation, elongation, and termination occur, and RNA could be maturated in nuclei. Additionally, non-coding RNAs (ncRNAs), including miRNAs, affect gene expression. Moreover, loop structures of DNAs also play roles in gene expression. One of the recent striking topics is the identification of extrachromosomal circular DNAs (eccDNAs), and R-loop formation that is mediated by the interaction between DNAs and RNAs. In summary, the regulation of gene expression is a very much complicated system. In this chapter, I would review how gene expression controlling systems in mammalian cells are presently understood. I hope that we would be inspired to think over essential problems to be dissolved toward the progress of medical sciences.

Advertisement

2. Gene expression, which is regulated by multiple steps

It has been widely known how general transcription factors (GTFs) execute RNA pol II dependent transcription in mammalian cells [1]. I have ever reviewed transcription control systems in mammalian cells, especially focusing on the possibility of the application of transcription-regulating mechanisms on gene therapy [2, 3]. This time, we would challenge the most fundamental problems that should be addressed before discussing the practical use of the transcription system. Generally, gene expression is defined as the producing rate of the mature mRNAs that are to be utilized for the translation of polypeptides. Transcription begins with initiation, elongation, and ends with the termination process. In eukaryotes, the premature RNAs are to be processed by splicing [4], 5′-cap modification [5, 6], and polyadenylation or poly A tail addition [7]. The matured RNAs are transported to the cytoplasm to be used for translation. RNAs are unstable and they are easily degraded by ribonucleases (RNases), which are ubiquitously present in all kinds of cells. In human, it has been known that at least 13 belong to hRNase A superfamily [8]. Some of them play roles in the host defense system [9], and others are required for host and mitochondrial DNA replication [10, 11, 12]. Thus, as same as other biologically synthesized polymers, amounts of matured RNAs are appropriately regulated by the balance between synthesis and degradation.

Advertisement

3. Loop Structures that regulate transcription

Generally, it is thought that gene expression is considerably regulated by the initiation step that is dependent on sequence-specific TFs [2, 3]. The loop structure formation by the interaction between enhancer and core promoter [13, 14, 15, 16] is thought to be essential for GTFs to start transcription in the right direction. The enhancer-promoter loop can be made by two double-stranded DNAs (dsDNA) and the most 5′-upstream RNA strand. The formation of chromatin loop clusters may be mediated by specific proteins, including CCCTC-binding factor (CTCF) [17]. The loop formation between dsDNAs might be associated with chromosome-wide spreading gene-silencing mechanism as that has been well studied for Xist [18]. The possible molecular model is that RNA strands may play a role as bridges between DNAs. On the other hand, R-loop structure, which enables the recycling of TFs and RNA pol II, is made by communication between transcription initiation and termination sites [19, 20, 21, 22]. The R-loop, which is composed of separated complementary DNA strands or “transcription bubble” and elongating RNA, could induce antisense transcription at enhancer elements, and transcription initiation and termination sites [23]. Thus, the DNA-RNA triplet R-loop can also play an essential role in the bidirectional transcription on eukaryotic chromosomes. The R-loop resembles to the replication bubble, which is made by separated complementary strands and a short RNA that is required for the leading strand synthesis. Therefore, the generation of R-loops needs to be accurately controlled. If the dysregulation occurred, it may lead to conflicts between transcription and replication machinery, which will cause DNA damage and cell death. The problem might be dissolved by several key regulators, including ATR, CHEK1, and BRD4 [24, 25]. The promoter regions of eukaryotic genes have been mainly studied by the reporter assay system with plasmid vectors that express such as luciferase. However, usually, it has not been paid attention to the 3′-untranslated regions (3′-UTRs). It has been presumed that 3′-UTRs play important roles in the regulation of stability, localization of transcripts, and translation [26, 27, 28]. If the R-loop has a considerable effect on the production of immature transcripts, it should not be ignored.

Advertisement

4. Transcriptional direction might be epigenetically regulated

Epigenetic regulation is mainly executed by DNA methylation and histone modification [29] that control chromatin structure to regulate genomic imprinting [30] and cellular senescence [31]. The direction of transcription might be dependent on DNA methylation, which is regulated by DNA methyltransferases (DNMTs) and the Ten-eleven-translocation (TET) enzymes [29, 32, 33]. CpG islands, which can be a methylation target [34], are commonly present at bidirectional transcription loci in human chromosomes [35, 36]. The bidirectional promoter regions have more GC-rich sequences but less TATA boxes than unidirectional promoters [37]. The majority (>80%) of CpGs in the human genome of the somatic cells are methylated, apart from actively transcribed regions, including promoters and enhancers [38]. However, because whole-genome methylome analyses identified differentially methylated regions (DMRs) in the human chromosomes [39], not all GC-rich sequences are the targets for methylation. Notably, specific TFs preferentially bind to the methylated CpGs [40]. ETS family protein PU.1 (SPI1) forms a complex with Dnmt3a/b to bring site-specific methylation to cause down-regulation of transcription [41]. The GC-box recognizing Sp1 can interact with DNMT1 in human cells [42]. Moreover, C/EBPα, Klf4, and Tfcp2l1 can affect Tet2 to demethylate specific promoters to induce pluripotency of cells [43]. In summary, site-specific DNA methylation/demethylation, modulating affinities with specific TFs, might determine which unidirectional or bidirectional transcription would be preferred.

Advertisement

5. Nutrients and metabolites dependent gene expression

Prokaryotic lactose operon system has been generally known. Nutrients or metabolites must be monitored to control transcription adequately in mammalian cells [44]. Glucose regulates the transcription of genes that encode lipogenesis-associated proteins through activation of the LXR (NR1H) factors [45]. HDL metabolism functioning protein-encoding genes are induced by glucose [46]. A glucose derivative molecule, 2-deoxy-D-glucose (2-DG) upregulates promoter activities of the TERT and WRN genes in HeLa S3 cells [47].

Fatty acids can affect transcription mediated by PPARs [48], SREBP-1 [49], and other TFs [50]. The n-butyrate (butyric acid), which is produced by gut bacteria, or sodium butyrate regulates gene expression in human cells [51]. The transcriptional regulation by butyrate has been explained by an inhibitory action on histone deacetylase [52] and increasing stabilities of mRNAs [53]. Notably, Sp1 that recognizes a GC-rich sequence [54] is hyper-acetylated by butyrate in human colon cells [55]. Therefore, transcription of specific genes, if their promoter contains a Sp1 binding element(s), could be affected by a lipophilic acid, which stops proliferation [56]. Other TFs, including ETS family ETV1 (ER81) [57], ETV4 (PEA3) [58], and p53 [59], are also activated by butyrate-induced signals.

Amino acids also regulate transcription. For example, glutamine responsive genes have been identified [60]. Leucine starvation induces promoter activity of the CHOP gene, encoding the CCAAT/enhancer binding protein, through binding of ATF-2 to an amino acid response element (AARE) [61]. Some amino acids are to be metabolized to acetyl-CoA and a methyl-group donor S-adenosylmethionine. They are the substrates for histone acetylation and DNA methylation, respectively. Tryptophan could be metabolized to a nicotinamide adenine dinucleotide (NAD+), which is also produced from niacin (nicotinic acid and nicotine amide) or vitamin B3, is not only required for mitochondrial functions, but also for poly (ADP-ribosyl)ation that is catalyzed by poly(ADP-ribose) polymerases (PARPs) [62, 63]. It has been suggested that NAD+ regulates transcription in mammalian cells [64, 65]. In mice, the hepatic and cytosolic NADH/NAD+ ratio alters the circulating α-hydroxybutyrate level, which could be an early biomarker for diabetes [66, 67]. Decrease in NAD+ concentration has been suggested to be associated with aging or age-related diseases [68]. Notably, 2-DG and trans-resveratrol, which up-regulate NAD+/NADH ratio in HeLa S3 cells [69], can activate the human WRN and TERT promoters [47, 70].

Advertisement

6. Euchromatin and heterochromatin

Euchromatin and heterochromatin represent specific structures of eukaryotic chromosomes, which are transcriptionally active and inactive, respectively [71]. Generally, heterochromatin represents a state where chromosomes are attached to a nuclear membrane with nuclear lamina [72]. The formation of the heterochromatin is thought to be under epigenetic regulation, during both the development and aging processes of mammalian cells. Heterochromatin does not only affect transcription but also protects chromosomes from mechanical stresses [73, 74]. Relationships between heterochromatin and DNA-repair systems have been proposed [75]. The DNA-repair systems could dominantly work on the euchromatin, and that is enabled by the stabilization of chromosomes by heterochromatin structure. It might be controversial that DNA repair factors cause heterochromatinization, but PARP-1 and 2, which play roles in the DNA-repair system, can contribute to the maintenance of heterochromatin [76]. The telomeric region where PARP or tankyrase is located plays a role in the maintenance of the ends of chromosomes [77]. The shortening of telomeres may lead to severe chromosomal instability that accelerates cellular senescence and cancer generation [78], suggesting that PARP enzymes protect telomeres by heterochromatinization. Moreover, the PARP modulates chromatin structure when it functions at centromeres [79]. Overall, PARP-dependent DNA repair systems are not only required for the conservation of nucleotide sequences of functional proteins but also for the maintenance of chromosomal structures that are constructed by specific or repetitive sequences. That would partly explain the reason why telomeres and centromeres might have excluded translocations of protein-encoding genes and transposons. This might also explain why PARP-1 does not have a preference for specific DNA sequences, surely it can find DNA breakage to load poly(ADP-ribose) to chromatin-associating proteins, including histones [80] and p53 [81]. The introduction of poly(ADP-ribose) on TFs may suppress transcription [82] and the repair system will work well in the promoter regions. Activation of PARPs consumes NAD+ to synthesize poly(ADP-ribose), which is required for indicating the DNA damaging sites. That will cause a reduction in NAD+-dependent transcription of mitochondrial protein-encoding genes [65]. Taken together, poly(ADP-ribosyl)ation plays a role in keeping a balance between DNA-repair, energy production, and transcription, maintaining chromosomal structures.

Advertisement

7. Concluding remarks

“Gene Expression” can be discussed from many points of view. Because it includes many biological events that are executed by various proteins and RNAs. Although I have not reviewed the recent progress in studies of non-coding RNAs, they play essential roles in transcriptional regulation [83]. “Gene Expression” is regulated by many stresses that can modulate DNA structures, loop formations, and epigenetic states. Dysregulation of “Gene Expression” will cause aging-related diseases. Hopefully, artificial transcription controlling systems will be developed and clinically applied to cancer and neurodegenerative diseases in the future. Nevertheless, we have not yet reached a conclusion or even a hypothesis on how gene expression system has been established and how it developed through a long evolution process. All organisms, including prokaryotes, archaea, and prokaryotes, would not live without accurate execution of the transcription system. The exceptions are viruses that just utilize the infected host cell system. Among them, some retroviruses are unique in carrying oncogenes to cause cancer and lymphoma [84]. Their genes are RNAs to be reverse transcribed to DNAs, which can be integrated into host cell chromosomes. The composition of the genome is characteristic, having long terminal repeats (LTRs). Interestingly, many retrovirus-like elements or transposons, including LINEs and SINEs, have been suggested to regulate gene expression, by both transcriptional and post-transcriptional mechanisms [85]. Every protein-encoding gene has a transcription start and termination site. How have genes acquired promoters and terminators, which are present on the 5′-upstreams and 3′-downstreams, respectively? The loop structures and extrachromosomal circular DNAs (eccDNAs), which are frequently identified when DNA amplification occurs, might give us a hint [86].

The dsDNA loops are thought to be formed when DNA damage was induced [87] or when a rearrangement of genes occurs in immune cells [88]. Although it is a hypothesis, the generation of multiple DNA replication initiation sites in eukaryotic chromosomes suggests that linear chromosomes might have been evolved from the fusion of multiple circular chromosomes (Figure 1). To prove this hypothesis, hot spot junction sites, which are the same as tentative dsDNA break sites, should be identified (Figure 1C). Loop or circular DNAs are formed at the time when dsDNAs integrated in or released from chromosomes. Therefore, there are both chances to gain or lose DNAs. If it occurred at chromosomal crossover during meiosis, the acquired or lost DNA sequences would be inherited to descendants. Elucidation of the biological meanings of transposition and amplification of genes will answer the question of how genes acquired promoters and terminators through evolution.

Figure 1.

DNA replication of eukaryotic cells. (A) Multiple DNA replication initiation sites (yellow circles) are distributed all over chromosomes. (B) Leading strands (red) and lagging strands (blue) are synthesized from the initiation sites when S-phase started. (C) At the sites where opposite direction proceeding leading strands collide, the temporary generated dsDNA breaks would be generated to be ligated. Taken together, multiple circular DNAs, which are to be ligated with, might be generated at the end of the S-phase.

References

  1. 1. Carey MF, Peterson CL, Smale ST. A primer on transcriptional regulation in mammalian cells. In: Carey MF, Peterson CL, Smale ST, editors. Transcriptional Regulation in Eukaryotes: Concepts, Strategies, and Techniques. 2nd ed. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press; 2009. pp. 1-45
  2. 2. Uchiumi F. Current studies in transcriptional control system; toward the establishment of therapies against human diseases. In: Uchiumi F, editor. Gene Expression and Regulation in Mammalian Cells. London, UK: InTechOPEN; 2018. pp. 3-13
  3. 3. Uchiumi F, Asai M. Gene expression controlling system and its application to medical sciences. In: Uchiumi F, editor. Gene Expression and Control. London, UK: InTechOPEN; 2019. pp. 3-10
  4. 4. Yan C, Wan R, Shi Y. Molecular mechanisms of pre-mRNA splicing through structural biology of the spliceosome. Cold Spring Harbor Perspectives in Biology. 2019;11(1):a032409
  5. 5. Muthukrishnan S, Both GW, Furuichi Y, Shatkin AJ. 5′-terminal 7-methylguanosine in eukaryotic mRNA is required for translation. Nature. 1975;255(5503):33-37
  6. 6. Sikorski PJ, Warminski M, Kubacka D, Ratajczak T, Nowis D, et al. The identity and methylation status of the first transcribed nucleotide in eukaryotic mRNA 5′ cap modulates protein expression in living cells. Nucleic Acids Research. 2020;48(4):1607-1626
  7. 7. Stewart M. Polyadenylation and nuclear export of mRNAs. The Journal of Biological Chemistry. 2019;294(9):2977-2987
  8. 8. Lee HH, Wang YN, Hung MC. Functional roles of the human ribonuclease A super family in RNA metabolism and membrane receptor biology. Molecular Aspects of Medicine. 2019;70:106-116
  9. 9. Rosenberg HF. RNase A ribonucleases and host defense: An evolving story. Journal of Leukocyte Biology. 2008;83(5):1079-1087
  10. 10. Parajuli S, Teasley DC, Murali B, Jackson J, Vindigni A, et al. Human ribonuclease H1 resolves R-loops and thereby enables progression of the DNA replication fork. The Journal of Biological Chemistry. 2017;292(37):15216-15226
  11. 11. Mentegari E, Crespan E, Bavagnoli L, Kissova M, Bertoletti F, et al. Ribonucleotide incorporation by human DNA polymerase eta impacts translesion synthesis and RNase H2 activity. Nucleic Acids Research. 2017;45(5):2600-2614
  12. 12. Posse V, Al-Behadili A, Uhler JP, Clausen AR, Reyes A, et al. RNase H1 directs origin-specific initiation of DNA replication in human mitochondria. PLoS Genetics. 2019;15(1):e1007781
  13. 13. van Arensbergen J, van Steensel B, Bussemaker HJ. In search of the determinants of enhancer-promoter interaction specificity. Trends in Cell Biology. 2014;24(11):695-702
  14. 14. Mora A, Sandve GK, Gabrielsen OS, Eskeland R. In the loop: Promoter-enhancer interactions and bioinformatics. Briefings in Bioinformatics. 2016;17(6):980-995
  15. 15. Robson MI, Ringel AR, Mundlos S. Regulatory landscaping: How enhancer-promoter communication is sculpted in 3D. Molecular Cell. 2019;74(6):1110-1122
  16. 16. Yokoshi M, Segawa K, Fukaya T. Visualizing the role of boundary elements in enhancer-promoter communication. Molecular Cell. 2020;78(2):224-235.e5
  17. 17. Li Y, Haarhuis JHI, Sedeño Cacciatore Á, Oldenkamp R, van Ruiten MS, et al. The structural basis for cohesin-CTCF-anchored loops. Nature. 2020;578(7795):472-476
  18. 18. Yamada N, Ogawa Y. Mechanisms of long noncoding Xist RNA-medated chromosome-wide gen silencing in X-chromosome inactivation. In: Kurokawa R, editor. Long Noncoding RNAs. Dordrecht Heidelberg London New York: Springer Science+Business Media; 2015. pp. 151-171
  19. 19. Martin M, Cho J, Cesare AJ, Griffith JD, Attardi G. Termination factor-mediated DNA loop between termination and initiation sites drives mitochondrial rRNA synthesis. Cell. 2005;123(7):1227-1240
  20. 20. El Kaderi B, Medler S, Raghunayakula S, Ansari A. Gene looping is conferred by activator-dependent interaction of transcription initiation and termination machineries. The Journal of Biological Chemistry. 2009;284(37):25015-25025
  21. 21. Bratkowski M, Unarta IC, Zhu L, Shubbar M, Huang X, et al. Structural dissection of an interaction between transcription initiation and termination factors implicated in promoter-terminator cross-talk. The Journal of Biological Chemistry. 2018;293(5):1651-1665
  22. 22. Malig M, Hartono SR, Giafaglione JM, Sanz LA, Chedin F. Ultra-deep coverage single-molecule R-loop footprinting reveals principles of R-loop formation. Journal of Molecular Biology. 2020;432(87):2271-2288
  23. 23. Tan-Wong SM, Dhir S, Proudfoot NJ. R-loop promote antisense transcription across the mammalian genome. Molecular Cell. 2019;76(4):600-616
  24. 24. Matos DA, Zhang JM, Ouyang J, Nguyen HD, Genois M, et al. ATR protects the genome against R loops through a MUS81-triggered feedback loop. Molecular Cell. 2019;77(3):514-527
  25. 25. Edwards DS, Maganti R, Tanksley JP, Luo J, Park JJH, et al. BRD4 prevents R-loop formation and transcription-replication conflicts by ensuring efficient transcription elongation. Cell Reports. 2020;32(12):108166
  26. 26. Geissler R, Grimson A. A position-specific 3′ UTR sequence that accelerates mRNA decay. RNA Biology. 2016;13(11):1075-1077
  27. 27. Mayr C. Regulation by 3′-untranslated regions. Annual Review of Genetics. 2017;51:171-194
  28. 28. Mayr C. What are 3′ UTRs doing? Cold Spring Harbor Perspectives in Biology. 2019;11(19):a034728
  29. 29. Loaeza-Loaeza J, Beltran AS, Hernández-Sotelo D. DNMTs and impact of CpG content, transcription factors, consensus motifs, lncRNAs, and histone marks on DNA methylation. Genes (Basel). 2020;11(11):1336
  30. 30. Meng H, Cao Y, Qin J, Song X, Zhang Q, et al. DNA methylation, its mediators and genome integrity. International Journal of Biological Sciences. 2015;11(5):604-617
  31. 31. Yang N, Sen P. The senescent cell epigenome. Aging (Albany NY). 2018;10(11):3590-3609
  32. 32. Ismail JN, Ghannam M, Al Outa A, Frey F, Shirinian M. Ten-eleven translocation proteins and their role beyond DNA demethylation—what we can learn from the fly. Epigenetics. 2020;15(11):1139-1150
  33. 33. Li E, Zhang Y. DNA methylation in mammals. Cold Spring Harbor Perspectives in Biology. 2014;6(5):a019133
  34. 34. Lakshminarasimhan R, Liang G. The role of DNA methylation in cancer. In: Jeltsch A, Jurkowska RZ, editors. DNA Methyltransferases-Role and Function. Advances in Experimental Medicine and Biology. Dordrecht Heidelberg London New York: Springer Science+Business Media; 2016. pp. 61-84
  35. 35. Adachi N, Lieber MR. Bidirectional gene organization: A common architectural feature of the human genome. Cell. 2002;109(7):807-809
  36. 36. Yang MQ, Elnitski L. Diversity of core promoter elements comprising human bidirectional promoters. BMC Genomics. 2008;9(Suppl. 2):S3
  37. 37. Trinklein ND, Aldred SF, Hartman SJ, Schroeder DI, Otillar RP, et al. An abundance of bidirectional promoters in the human genome. Genome Research. 2004;14(1):62-66
  38. 38. Luo C, Hajkova P, Ecker JR. Dynamic DNA methylation: In the right place at the right time. Science. 2018;361(6409):1336-1340
  39. 39. Docherty LE, Rezwan FI, Poole RL, Jagoe H, Lake H, et al. Genome-wide DNA methylation analysis of patients with imprinting disorders identifies differentially methylated regions associated with novel candidate imprinted genes. Journal of Medical Genetics. 2014;51(4):229-238
  40. 40. Yin Y, Morgunova E, Jolma A, Kaasinen E, Sahu B, et al. Impact of cytosine methylation on DNA binding specificities of human transcription factors. Science. 2017;356(6337):eaaj2239
  41. 41. Suzuki M, Yamada T, Kihara-Negishi F, Sakurai T, Hara E, et al. Site-specific DNA methylation by a complex of PU.1 and Dnmt3a/b. Oncogene. 2006;15(17):2477-2488
  42. 42. Hervouet E, Vallette FM, Cartron PF. Dnmt1/transcription factor interactions: An alternative mechanism of DNA methylation inheritance. Genes & Cancer. 2010;1(5):434-443
  43. 43. Sardina JL, Collombet S, Tian TV, Gómez A, Di Stefano B, et al. Transcription factors drive Tet2-mediated enhancer demethylation to reprogram cell fate. Cell Stem Cell. 2018;23(5):727-741
  44. 44. Haro D, Marrero PF, Relat J. Nutritional regulation of gene expression: Carbohydrate-, Fat-, and acmino acid-dependent modulation of transcriptional activity. International Journal of Molecular Sciences. 2019;20(6):1386
  45. 45. Mitro N, Mak PA, Vargas L, Godio C, Hampton E, et al. The nuclear receptor LXR is a glucose sensor. Nature. 2007;445(7124):219-223
  46. 46. Tu AY, Albers JJ. Glucose regulates the transcription of human genes relevant to HDL metabolism: Responsive elements for peroxisome proliferator-activated receptor are involved in the regulation of phospholipid transfer protein. Diabetes. 2001;50(8):1851-1856
  47. 47. Zhou B, Ikejima T, Watanabe T, Iwakoshi K, Idei Y, et al. The effect of 2-deoxy-D-glucose on Werner syndrome RecQ helicase gene. FEBS Letters. 2009;583(8):1331-1336
  48. 48. Latruffe N, Cherkaoui Malki M, Nicolas-Frances V, Clemencet MC, et al. Regulation of the peroxisomal β-oxidation-dependent pathway by peroxisome proliferator-activated receptor α and kinases. Biochemical Pharmacology. 2000;60(8):1027-1032
  49. 49. Magaña MM, Osborne TF. Two tandem binding sites for sterol regulatory element binding proteins are required for sterol regulation of fatty-acid synthase promoter. The Journal of Biological Chemistry. 1996;271(51):32689-32694
  50. 50. Jump DB, Tripathy S, Depner CM. Fatty acid-regulated transcription factors in the liver. Annual Review of Nutrition. 2013;33:249-269
  51. 51. Vanhoutvin SA, Troost FJ, Hamer HM, Lindsey PJ, Koek GH, et al. Butyrate-induced transcriptional changes in human colonic mucosa. PLoS One. 2009;4(8):e6759
  52. 52. Chriett S, Dąbek A, Wojtala M, Vidal H, Balcerczyk A, et al. Prominent action of butyrate over beta-hydroxybutyrate as histone deacetylase inhibitor, transcriptional modulator and anti-inflammatory molecule. Scientific Reports. 2019;9(1):742
  53. 53. Torun A, Enayat S, Sheraj I, Tunçer S, Ülgen DH, et al. Butyrate mediated regulation of RNA binding proteins in the post-transcriptional regulation of inflammatory gene expression. Cellular Signalling. 2019;64:109410
  54. 54. Wierstra I. Sp1: Emerging roles - beyond constitutive activation of TATA-less housekeeping genes. Biochemical and Biophysical Research Communications. 2008;372(1):1-13
  55. 55. Waby JS, Chirakkal H, Yu C, Griffiths GJ, Benson RS, et al. Sp1 acetylation is associated with loss of DNA binding at promoters associated with cell cycle arrest and cell death in a colon cell line. Molecular Cancer. 2010;15(9):275
  56. 56. Ginsburg E, Salomon D, Sreevalsan T, Freese E. Growth inhibition and morphological change caused by lipophilic acids in mammalian cells. Proceedings of the National Academy of Sciences of the United States of America. 1973;70(8):2457-2461
  57. 57. Goel A, Janknecht R. Acetylation-mediated transcriptional activation of the ETS protein ER81 by p300, P/CAF, and HER2/Neu. Molecular and Cellular Biology. 2003;23(17):6243-6254
  58. 58. Guo B, Panagiotaki N, Warwood S, Sharrocks AD. Dynamic modification of the ETS transcription factor PEA3 by sumoylation and p300-mediated acetylation. Nucleic Acids Research. 2011;39(15):6403-6413
  59. 59. Gaub P, Tedeschi A, Puttagunta R, Nguyen T, Schmandke A, et al. HDAC inhibition promotes neural outgrowth and counteracts growth cone collapse through CBP/p300 and P/CAF-dependent p53 acetylation. Cell Death and Differentiation. 2010;17(9):1392-1408
  60. 60. Brasse-Lagnel C, Lavoinne A, Husson A. Control of mammalian gene expression by amino acids, especially glutamine. The FEBS Journal. 2009;276(7):1826-1844
  61. 61. Bruhat A, Jousse C, Carraro V, Reimold AM, Ferrara M, et al. Amino acids control mammalian gene transcription: Activating transcription factor 2 is essential for the amino acid responsiveness of the CHOP promoter. Molecular and Cellular Biology. 2000;20(19):7192-7240
  62. 62. Bender DA. Micronutrients. In: Bender DA, editor. Introduction to Nutrition and Metabolism. 5th ed. Boca Raton, FL: CRC Press, Taylor & Francis Group, Inc.; 2014. pp. 307-384
  63. 63. Uchiumi F, Sato A, Asai M, Tanuma S. An NAD+ dependent/sensitive transcription system: Toward a novel anti-cancer therapy. AIMS Molecular Science. 2020;7(1):12-28
  64. 64. Ghosh S, George S, Roy U, Ramachandran D, Kolthur-Seetharam U. NAD: A master regulator of transcription. Biochimica et Biophysica Acta. 2010;1799(10-12):681-693
  65. 65. Ryu KW, Nandu T, Kim J, Challa S, DeBerardinis RJ, et al. Metabolic regulation of transcription through compartmentalized NAD+ biosynthesis. Science. 2018;360(6389):eaan5780
  66. 66. Goodman RP, Markhard AL, Shah H, Sharma R, Skinner OS, et al. Hepatic NADH reductive stress underlies common variation in metabolic traits. Nature. 2020;583(7814):122-126
  67. 67. Gall WE, Beebe K, Lawton KA, Adam KP, Mitchell MW, et al. α-Hydroxybutyrate is an early biomarker of insulin resistance and glucose intolerance in a nondiabetic population. PLoS One. 2010;5(5):e10883
  68. 68. Rehmani I, Liu F, Liu A. Cell signaling and transcription. In: Villamena FA, editor. Molecular Basis of Oxidative Stress: Chemistry, Mechanisms, and Disease Pathogenesis. Hoboken, NJ: John Wiley & Sons; 2013. pp. 179-201
  69. 69. Takihara Y, Sudo D, Arakawa J, Takahashi M, Sato A, et al. Nicotinamide adenine dinucleotide (NAD+) and cell aging. In: Strakoš R, Lorens B, editors. New Research on Cell Aging and Death. Hauppauge, NY: Nova Science Publishers, Inc.; 2018. pp. 131-158
  70. 70. Uchiumi F, Watanabe T, Hasegawa S, Hoshi T, Higami Y, et al. The effect of Resveratrol on the Werner Syndrome RecQ helicase gene and telomerase activity. Current Aging Science. 2011;4:1-7
  71. 71. Liu J, Ali M, Zhou Q. Establishment and evolution of heterochromatin. Annals of the New York Academy of Sciences. 2020;1476(1):59-77
  72. 72. van Steensel B, Belmont AS. Lamina-associated domains: Link with chromosome architecture, heterochromatin, and gene repression. Cell. 2017;169(5):780-791
  73. 73. Dillon N. Heterochromatin structure and function. Biology of the Cell. 2004;96(8):631-637
  74. 74. Nava MM, Miroshnikova YA, Biggs LC, Whitefield DB, Metge F, et al. Heterochromatin-driven nuclear softening protects the genome against mechanical stress-induced damage. Cell. 2020;181(4):800-817
  75. 75. Fortuny A, Polo SE. The response to DNA damage in heterochromatin domains. Chromosoma. 2018;127(3):291-300
  76. 76. Dantzer F, Santoro R. The expanding role of PARPs in the establishment and maintenance of heterochromatin. The FEBS Journal. 2013;280(15):3508-3518
  77. 77. Smith S, Giriat I, Schmitt A, de Lange T. Tankyrase, a poly(ADP-ribose)polymerase at human telomeres. Science. 1998;282(5393):1484-1487
  78. 78. Mathieu N, Pirzio L, Freulet-Marrière MA, Desmaze C, Sabatier L. Telomeres and chromosomal instability. Cellular and Molecular Life Sciences. 2004;61(6):641-656
  79. 79. Earle E, Saxena A, MacDonald A, Hudson DF, Shaffer LG, et al. Poly(ADP-ribose) polymerase at active centromeres and neocentromeres at metaphase. Human Molecular Genetics. 2000;9(2):187-194
  80. 80. Naegeli H, Althaus FR. Regulation of poly(ADP-ribose) polymerase. Histone-specific adaptations of reaction products. The Journal of Biological Chemistry. 1991;266(16):10596-10601
  81. 81. Kim HL, Ra H, Kim KR, Lee JM, Im H, et al. Poly(ADP-ribosyl)ation of p53 contributes to TPEN-induced neuronal apoptosis. Molecules and Cells. 2015;38(4):312-317
  82. 82. Ding L, Chen X, Xu X, Qian Y, Liang G, et al. PARP1 suppresses the transcription of PD-L1 by poly(ADP-ribosyl)ating STAT3. Cancer Immunology Research. 2019;7(1):136-149
  83. 83. Dykes IM, Emanueli C. Transcriptional and post-transcriptional gene regulation by long non-coding RNA. Genomics, Proteomics & Bioinformatics. 2017;15(3):177-186
  84. 84. Bishop JM. Enemies within: The genesis of retrovirus oncogenes. Cell. 1981;23(1):5-6
  85. 85. Elbarbary RA, Lucas BA, Maquat LE. Retrotransposons as regulators of gene expression. Science. 2016;351(6274):aac7247
  86. 86. Ling X, Han Y, Meng J, Zhong B, Chen J, et al. Small extrachromosomal circular DNA (eccDNA): Major functions in evolution and cancer. Molecular Cancer. 2021;20(1):113
  87. 87. Paulsen T, Kumar P, Koseoglu MM, Dutta A. New discoveries of extrachromosomal circles of DNA in normal and tumor cells. Trends in Genetics. 2018;34(4):270-278
  88. 88. Peters JM. How DNA loop extrusion mediated by cohesin enables V(D)J recombination. Current Opinion in Cell Biology. 2021;71:75-83

Written By

Fumiaki Uchiumi

Submitted: 15 January 2022 Published: 05 October 2022