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 -- 1490 2023-04-19 06:17:38 |
2 format corrected. + 2 word(s) 1492 2023-04-19 07:09:46 |

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.
Elson, D.J.; Kolluri, S.K. Aryl Hydrocarbon Receptor Biology and Signaling. Encyclopedia. Available online: https://encyclopedia.pub/entry/43220 (accessed on 27 July 2024).
Elson DJ, Kolluri SK. Aryl Hydrocarbon Receptor Biology and Signaling. Encyclopedia. Available at: https://encyclopedia.pub/entry/43220. Accessed July 27, 2024.
Elson, Daniel J., Siva K. Kolluri. "Aryl Hydrocarbon Receptor Biology and Signaling" Encyclopedia, https://encyclopedia.pub/entry/43220 (accessed July 27, 2024).
Elson, D.J., & Kolluri, S.K. (2023, April 19). Aryl Hydrocarbon Receptor Biology and Signaling. In Encyclopedia. https://encyclopedia.pub/entry/43220
Elson, Daniel J. and Siva K. Kolluri. "Aryl Hydrocarbon Receptor Biology and Signaling." Encyclopedia. Web. 19 April, 2023.
Aryl Hydrocarbon Receptor Biology and Signaling
Edit

The aryl hydrocarbon receptor (AhR) is a ligand-activated transcription factor involved in regulating a wide range of biological responses. A diverse array of xenobiotics and endogenous small molecules bind to the receptor and drive unique phenotypic responses. Due in part to its role in mediating toxic responses to environmental pollutants, AhR activation has not been traditionally viewed as a viable therapeutic approach. Nonetheless, the expression and activation of AhR can inhibit the proliferation, migration, and survival of cancer cells, and many clinically approved drugs transcriptionally activate AhR. Identification of novel select modulators of AhR-regulated transcription that promote tumor suppression is an active area of investigation. The development of AhR-targeted anticancer agents requires a thorough understanding of the molecular mechanisms driving tumor suppression.

aryl hydrocarbon receptor endogenous differentiation co-regulators

1. Role of AhR in Xenobiotic Metabolism

AhR belongs to the superfamily of transcription factors containing both basic helix–loop–helix (bHLH) and PER-ARNT-SIM (PAS) domains (14). The PAS domain is conserved across all kingdoms of life, and functions in sensing and activating biological responses to environmental changes. Related transcription factors containing the PAS domain, such as hypoxia-inducible factor 1α (HIF1α), and circadian clock proteins such as PER1 and BMAL, respond to environmental cues such as oxygen tension and light [1]. In this respect, the aryl hydrocarbon receptor functions as both an internal and external sensor of chemical signals.
The role of AhR in the cellular detoxification or metabolism of small molecules is well established [2]. Exposure to xenobiotics, or activation by endogenous metabolites, induces the receptor’s activation and a subsequent transcriptional response that then attempts to bring the cell back to a state of homeostasis. A major component of this homeostatic response is induction of the cytochrome P450 enzymes, including CYP1A1, CYP1A2, and CYP1B1, which function, through substrate oxidation, to convert the ligands to more water-soluble forms that can be conjugated by Phase II enzymes (e.g., glutathione-S-transferase) and excreted via Phase III enzymatic transport [3]. While it was evolutionarily acquired as a transcriptional circuit to ameliorate xenobiotic insults, exposure to certain xenobiotics can induce AhR-mediated toxicity. It is in this context that AhR has been historically and extensively studied, as AhR was originally cloned and identified as the receptor responsible for mediating the biological effects of dioxin (2,3,7,8-tetrachloro-dibenzo-p-dioxin (TCDD)) [4], a legacy anthropogenic pollutant. TCDD binds with high-affinity to AhR, is resistant to CYP1-mediated metabolic degradation, and is highly lipophilic. Together, these properties make TCDD highly bio-accumulative, and its toxic effects are from chronic AhR activation [5].
Within an organism, activation and antagonism of AhR have diverse consequences that depend critically on the cellular abundance and characteristics of the initiating ligand, as well as specific tissue and organ dynamics. While AhR is a master regulator of xenobiotic metabolism, it also has important roles in processes such as development, differentiation, and immune function.

2. The AhR Signaling Pathway

The transcriptionally inactive AhR complex is located in the cytosol, sequestered by an HSP90 dimer [6][7], co-chaperone p23 [8][9], AhR-interacting protein (AIP), and the protein kinase Src [10]. The complex functions to both sequester AhR in the cytosol and hold the AhR in conformation to interact with ligands [11]. Upon an agonistic ligand binding to the receptor, AhR-associated AIP dissociates from the receptor, exposing its nuclear localization sequence, and importin-β then binds to AhR and transports the receptor into the nucleus.
Once in the nucleus, AhR can then interact with the heterodimer partner, aryl hydrocarbon receptor nuclear translocator (Arnt), also known as HIF-1β, where the AhR–Arnt complex then binds to the regulatory regions of DNA containing the consensus motif (5′-TNGCGTG-3′) [12][13] or the xenobiotic response element (XRE) (also known as dioxin response element (DRE), or the AhR response element (AhRE) and modulates the transcription of the target gene. While XRE-driven transcription through AhR–Arnt signaling represents the most well studied mode of regulation, AhR can modulate gene expression through non-XRE elements [14][15] and can also interact with other transcription factors including c- MAF [16], KLF6 [17], RelA [18], and other NF-κB complex members [19].
Structurally, AhR contains the following features, proceeding from its amino terminus: a bHLH domain (amino acids (a.a.) 33–87); PAS-A (a.a. 111–273) and PAS-B domains (a.a. 275–386); and a transactivation domain (TAD) that contains an acidic region, a glutamine-rich (Q-rich) region, and a proline/serine-rich region (a.a. 490–805) [20][21]. AhR binds to the regulatory regions of DNA via its bHLH domain [20], which is masked by one molecule of HSP90 in the unliganded cytoplasmic complex [6], partially overlapping with the PAS-A region. The second molecule of HSP90 interacts dominantly with the PAS-B region. The bHLH domain dimerizes with the bHLH of Arnt to form a bundle of four helixes that interact with the XRE [12]. Similarly, the PAS-A domains of AhR and Arnt function as a dimerization interface, conferring specificity and enhancing the complex’s stability and DNA binding [22][23]. The PAS-B domain functions as the ligand-binding domain [23] and is required for nuclear translocation of the receptor, but is dispensable for heterodimerization and ligand-binding [24][25]. Notably, high-affinity ligand-binding and the subsequent induction of metabolizing enzymes represents a vertebrate adaptation and has been suggested as an acquired function in response to aromatic marine natural products [26]. Early invertebrate AhR paralogs in Caenhorabiditis elegans [27][28] and Drosophila melanogaster [29] function primarily in neuronal and sensory development, and this role is retained in vertebrates in addition to its ligand-mediated activities [30].
Species-specific [31], ligand-specific [32][33], dose-dependent [34], tissue-specific, and microenvironment-specific factors [35][36] together influence the apical biological outcome downstream of the activation of the AhR ligand, so making initial predictions of a ligand’s effects in a given model system is extremely difficult. For example, species-specific divergences in the receptor’s structure and ligand-binding affinity result in significant differences in the transcriptional [31] and phenotypic response to ligands such as TCDD between humans [31], mice [37], and guinea pigs [38]. The mouse AhR differs from its human ortholog in its ligand-binding affinity due to a single amino acid substitution in the ligand-binding pocket [39]. Consequently, TCDD binds with significantly greater affinity (approximately 10×) to the mouse AhR, while distinct ligands such as indoxyl-3-sulfate (of endogenous origin) bind with greater affinity to the human receptor [39][40]. To understand some of these interspecies differences, the researchers' laboratory previously generated an AhR homology model for the AhR PAS domain [41][42].

3. Role of Co-Regulators in AhR Signaling

The presence or absence of co-regulatory proteins and their activity levels influence both the basal transcription and ligand inducibility of AhR [43][44] (Figure 1). Recently, hexokinase 2 was found to be a transcriptional target of AhR and a positive regulator of AhR-mediated transcription [45]. Important regulators of AhR signaling also include proteins such as AhRR [46], TiPARP [47][48][49][50], Gadd45b [51], p300, CREB binding protein (p300/CBP) [52], Smad3 [53], and SIN3A [54][55]. AhRR and TiPARP function to provide negative feedback on AhR’s transcriptional activity, while SIN3A, typically regarded as a transcriptional repressor, was found to function as an essential co-activator for the transcription of CYP1A1 [55]. An additional mechanism fine-tuning the function of AhR was recently identified [56]. Bourner et al. found that differential expression levels of ARNT isoforms altered the transcriptional activity of AhR, and ARNT isoform-dependent regulation of AhR was shown to be dependent on phosphorylation by the AhR target gene casein kinase 2 [56].
Figure 1. Summary of the positive and negative transcriptional co-regulators of AhR signaling. The diagram depicts AhR bound to DNA in complex with potential heterodimer partner proteins and regulators. Different positive and negative regulators of AhR-mediated transcription are summarized above.
The researchers recently discovered that the expression of the cell cycle inhibitor and direct AhR target gene, p27Kip1, represses AhR-mediated transcription [57] (Figure 2). Loss of p27Kip1 in lung cancer cells resulted in significantly enhanced basal and ligand-induced levels of AhR target genes such as CYP1A1 and AHRR, and this negative regulation was found for the transcription of p27Kip1 itself [57]. Interestingly, p27Kip1-dependent transcriptional repression was previously found to operate via the formation of a protein complex containing mouse retinoblastoma protein (Rb) homolog p130, mSIN3A, E2F4, and multiple histone deacetylases (HDACs 1, 4, and 5) [58][59], Further, AhR’s interaction with Rb [60] and transcriptional regulation of AhR downstream from Rb have been reported [61], indicating bidirectional regulation between AhR and components of the cell cycle’s machinery.
Figure 2. Negative regulation of AhR-mediated transcription by the expression of p27Kip1. The diagram depicts the regulatory relationship between p27Kip1 and AhR, whereby AhR acts as a direct transcriptional regulator of the transcription of p27Kip1, and increased expression of p27Kip1 both inhibits cell cycle progression and represses the AhR-mediated transcription of genes such as CDKN1A (encoding p21Cip1), AHRR, and CDKN1B (encoding p27Kip1 itself) as an auto-feedback loop.
The p300/CBP enzymes function as transcriptional co-activators via their acetyltransferase activity, and the catalytic activity of p300/CBP was found to be essential for AhR’s transcriptional activity [52]. Notably, ARNT was identified as the key target of p300/CBP’s acetylation for activating transcription [52]. Interestingly, previous reports indicated that the acetylation of the p27Kip1 by p300/CBP-associated factor (PCAF) attenuated its transcriptional repressor activity and promoted its degradation, suggesting an additional layer of regulatory control [62][63].

References

  1. McIntosh, B.E.; Hogenesch, J.B.; Bradfield, C.A. Mammalian Per-Arnt-Sim Proteins in Environmental Adaptation. Annu. Rev. Physiol. 2010, 72, 625–645.
  2. Denison, M.S.; Soshilov, A.A.; He, G.; DeGroot, D.E.; Bin Zhao, B. Exactly the Same but Different: Promiscuity and Diversity in the Molecular Mechanisms of Action of the Aryl Hydrocarbon (Dioxin) Receptor. Toxicol. Sci. 2011, 124, 1–22.
  3. Ciolino, H.P.; Macdonald, C.J.; Memon, O.S.; Bass, S.E.; Yeh, G.C. Sulindac regulates the aryl hydrocarbon receptor-mediated expression of Phase 1 metabolic enzymes in vivo and in vitro. Carcinogenesis 2005, 27, 1586–1592.
  4. Poland, A.P.; Glover, E.; Robinson, J.R.; Nebert, D.W. Genetic expression of aryl hydrocarbon hydroxylase activity. Induction of monooxygenase activities and cytochrome P1-450 formation by 2,3,7,8 tetrachlorodibenzo p dioxin in mice genetically “nonresponsive” to other aromatic hydrocarbons. J. Biol. Chem. 1974, 249, 5599–5606.
  5. National Academies of Sciences, Engineering, and Medicine. Veterans and Agent Orange: Update 2014; The National Academies Press: Washington, DC, USA, 2016.
  6. Antonsson, C.; Whitelaw, M.L.; McGuire, J.; Gustafsson, J.A.; Poellinger, L. Distinct roles of the molecular chaperone hsp90 in modulating dioxin receptor function via the basic helix-loop-helix and PAS domains. Mol. Cell. Biol. 1995, 15, 756–765.
  7. Tsuji, N.; Fukuda, K.; Nagata, Y.; Okada, H.; Haga, A.; Hatakeyama, S.; Yoshida, S.; Okamoto, T.; Hosaka, M.; Sekine, K.; et al. The activation mechanism of the aryl hydrocarbon receptor (AhR) by molecular chaperone HSP90. FEBS Open Bio 2014, 4, 796–803.
  8. Pappas, B.; Yang, Y.; Wang, Y.; Kim, K.; Chung, H.J.; Cheung, M.; Ngo, K.; Shinn, A.; Chan, W.K. p23 protects the human aryl hydrocarbon receptor from degradation via a heat shock protein 90-independent mechanism. Biochem. Pharmacol. 2018, 152, 34–44.
  9. Meyer, B.K.; Pray-Grant, M.G.; Heuvel, J.P.V.; Perdew, G.H. Hepatitis B Virus X-Associated Protein 2 Is a Subunit of the Unliganded Aryl Hydrocarbon Receptor Core Complex and Exhibits Transcriptional Enhancer Activity. Mol. Cell. Biol. 1998, 18, 978–988.
  10. Enan, E.; Matsumura, F. Identification of c-Src as the integral component of the cytosolic Ah receptor complex, transducing the signal of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) through the protein phosphorylation pathway. Biochem. Pharmacol. 1996, 52, 1599–1612.
  11. Petrulis, J.R.; Perdew, G.H. The role of chaperone proteins in the aryl hydrocarbon receptor core complex. Chem. Biol. Interact. 2002, 141, 25–40.
  12. Yao, E.F.; Denison, M.S. DNA sequence determinants for binding of transformed Ah receptor to a dioxin-responsive enhancer. Biochemistry 1992, 31, 5060–5067.
  13. Swanson, H.; Tullis, K.; Denison, M.S. Binding of transformed Ah receptor complex to a dioxin responsive transcriptional enhancer: Evidence for two distinct heteromeric DNA-binding forms. Biochemistry 1993, 32, 12841–12849.
  14. Tanos, R.; Patel, R.D.; Murray, I.A.; Smith, P.B.; Patterson, A.D.; Perdew, G.H. Aryl hydrocarbon receptor regulates the cholesterol biosynthetic pathway in a dioxin response element-independent manner. Hepatology 2012, 55, 1994–2004.
  15. Jackson, D.P.; Li, H.; Mitchell, K.A.; Joshi, A.D.; Elferink, C.J. Ah Receptor–Mediated Suppression of Liver Regeneration through NC-XRE–Driven p21Cip1Expression. Mol. Pharmacol. 2014, 85, 533–541.
  16. Apetoh, L.; Quintana, F.J.; Pot, C.; Joller, N.; Xiao, S.; Kumar, D.; Burns, E.J.; Sherr, D.H.; Weiner, H.L.; Kuchroo, V.K. The aryl hydrocarbon receptor interacts with c-Maf to promote the differentiation of type 1 regulatory T cells induced by IL-27. Nat. Immunol. 2010, 11, 854–861.
  17. Wilson, S.R.; Joshi, A.D.; Elferink, C.J. The Tumor Suppressor Kruppel-Like Factor 6 Is a Novel Aryl Hydrocarbon Receptor DNA Binding Partner. J. Pharmacol. Exp. Ther. 2013, 345, 419–429.
  18. Kim, D.W.; Gazourian, L.; Quadri, S.A.; Romieu-Mourez, R.; Sherr, D.H.; Sonenshein, G.E. The RelA NF-kappaB subunit and the aryl hydrocarbon receptor (AhR) cooperate to transactivate the c-myc promoter in mammary cells. Oncogene 2000, 19, 5498–5506.
  19. Vogel, C.F.A.; Matsumura, F. A new cross-talk between the aryl hydrocarbon receptor and RelB, a member of the NF-κB family. Biochem. Pharmacol. 2009, 77, 734–745.
  20. Reisz-Porszasz, S.; Probst, M.R.; Fukunaga, B.N.; Hankinson, O. Identification of functional domains of the aryl hydrocarbon receptor nuclear translocator protein (ARNT). Mol. Cell. Biol. 1994, 14, 6075–6086.
  21. Crews, S.T.; Fan, C.-M. Remembrance of things PAS: Regulation of development by bHLH–PAS proteins. Curr. Opin. Genet. Dev. 1999, 9, 580–587.
  22. Lindebro, M.; Poellinger, L.; Whitelaw, M. Protein-protein interaction via PAS domains: Role of the PAS domain in positive and negative regulation of the bHLH/PAS dioxin receptor-Arnt transcription factor complex. EMBO J. 1995, 14, 3528–3539.
  23. Pongratz, I.; Antonsson, C.; Whitelaw, M.L.; Poellinger, L. Role of the PAS Domain in Regulation of Dimerization and DNA Binding Specificity of the Dioxin Receptor. Mol. Cell. Biol. 1998, 18, 4079–4088.
  24. Andersson, P.; McGuire, J.; Rubio, C.; Gradin, K.; Whitelaw, M.L.; Pettersson, S.; Hanberg, A.; Poellinger, L. A constitutively active dioxin/aryl hydrocarbon receptor induces stomach tumors. Proc. Natl. Acad. Sci. USA 2002, 99, 9990–9995.
  25. Schulte, K.W.; Green, E.; Wilz, A.; Platten, M.; Daumke, O. Structural Basis for Aryl Hydrocarbon Receptor-Mediated Gene Activation. Structure 2017, 25, 1025–1033.e3.
  26. Hahn, M.E.; Karchner, S.I.; Merson, R.R. Diversity as opportunity: Insights from 600 million years of AHR evolution. Curr. Opin. Toxicol. 2017, 2, 58–71.
  27. Huang, X.; Powell-Coffman, J.A.; Jin, Y. The AHR-1 aryl hydrocarbon receptor and its co-factor the AHA-1 aryl hydrocarbon receptor nuclear translocator specify GABAergic neuron cell fate in C. elegans. Development 2004, 131, 819–828.
  28. Qin, H.; Powell-Coffman, J.A. The Caenorhabditis elegans aryl hydrocarbon receptor, AHR-1, regulates neuronal development. Dev. Biol. 2004, 270, 64–75.
  29. Sonnenfeld, M.; Ward, M.; Nystrom, G.; Mosher, J.; Stahl, S.; Crews, S. The Drosophila tango gene encodes a bHLH-PAS protein that is orthologous to mammalian Arnt and controls CNS midline and tracheal development. Development 1997, 124, 4571–4582.
  30. Di Giaimo, R.; Durovic, T.; Barquin, P.; Kociaj, A.; Lepko, T.; Aschenbroich, S.; Breunig, C.T.; Irmler, M.; Cernilogar, F.M.; Schotta, G.; et al. The Aryl Hydrocarbon Receptor Pathway Defines the Time Frame for Restorative Neurogenesis. Cell Rep. 2018, 25, 3241–3251.e5.
  31. Rothhammer, V.; Borucki, D.M.; Kenison, J.E.; Hewson, P.; Wang, Z.; Bakshi, R.; Sherr, D.H.; Quintana, F.J. Detection of aryl hydrocarbon receptor agonists in human samples. Sci. Rep. 2018, 8, 4970.
  32. Wheeler, J.L.H.; Martin, K.C.; Resseguie, E.; Lawrence, B.P. Differential Consequences of Two Distinct AhR Ligands on Innate and Adaptive Immune Responses to Influenza A Virus. Toxicol. Sci. 2014, 137, 324–334.
  33. Bohonowych, J.E.; Denison, M.S. Persistent Binding of Ligands to the Aryl Hydrocarbon Receptor. Toxicol. Sci. 2007, 98, 99–109.
  34. Ehrlich, A.K.; Pennington, J.M.; Bisson, W.H.; Kolluri, S.K.; Kerkvliet, N.I. TCDD, FICZ, and Other High Affinity AhR Ligands Dose-Dependently Determine the Fate of CD4+ T Cell Differentiation. Toxicol. Sci. 2018, 161, 310–320.
  35. Vogel, C.F.; Khan, E.M.; Leung, P.S.; Gershwin, M.E.; Chang, W.L.; Wu, D.; Haarmann-Stemmann, T.; Hoffmann, A.; Denison, M.S. Cross-talk between Aryl Hydrocarbon Receptor and the Inflammatory Response: A role for nuclear factor-κB. J. Biol. Chem. 2014, 289, 1866–1875.
  36. DiNatale, B.C.; Smith, K.; John, K.; Krishnegowda, G.; Amin, S.G.; Perdew, G.H. Ah Receptor Antagonism Represses Head and Neck Tumor Cell Aggressive Phenotype. Mol. Cancer Res. 2012, 10, 1369–1379.
  37. Fletcher, N.; Hanberg, A.; Håkansson, H. Hepatic Vitamin A Depletion Is a Sensitive Marker of 2,3,7,8-Tetrachlorodibenzo-p-dioxin (TCDD) Exposure in Four Rodent Species. Toxicol. Sci. 2001, 62, 166–175.
  38. Korkalainen, M.; Tuomisto, J.; Pohjanvirta, R. The AH Receptor of the Most Dioxin-Sensitive Species, Guinea Pig, Is Highly Homologous to the Human AH Receptor. Biochem. Biophys. Res. Commun. 2001, 285, 1121–1129.
  39. Ramadoss, P.; Perdew, G.H. Use of 2-Azido-3-iodo-7,8-dibromodibenzo-p-dioxin as a Probe to Determine the Relative Ligand Affinity of Human versus Mouse Aryl Hydrocarbon Receptor in Cultured Cells. Mol. Pharmacol. 2004, 66, 129–136.
  40. Hubbard, T.D.; Murray, I.A.; Perdew, G.H. Indole and Tryptophan Metabolism: Endogenous and Dietary Routes to Ah Receptor Activation. Drug Metab. Dispos. 2015, 43, 1522–1535.
  41. Perkins, A.; Phillips, J.L.; Kerkvliet, N.I.; Tanguay, R.L.; Perdew, G.H.; Kolluri, S.K.; Bisson, W.H. A Structural Switch between Agonist and Antagonist Bound Conformations for a Ligand-Optimized Model of the Human Aryl Hydrocarbon Receptor Ligand Binding Domain. Biology 2014, 3, 645–669.
  42. Bisson, W.H.; Koch, D.C.; O’Donnell, E.F.; Khalil, S.M.; Kerkvliet, N.I.; Tanguay, R.L.; Abagyan, R.; Kolluri, S.K. Modeling of the Aryl Hydrocarbon Receptor (AhR) Ligand Binding Domain and Its Utility in Virtual Ligand Screening to Predict New AhR Ligands. J. Med. Chem. 2009, 52, 5635–5641.
  43. Wang, S.; Hankinson, O. Functional Involvement of the Brahma/SWI2-related Gene 1 Protein in Cytochrome P4501A1 Transcription Mediated by the Aryl Hydrocarbon Receptor Complex. J. Biol. Chem. 2002, 277, 11821–11827.
  44. Hankinson, O. Role of coactivators in transcriptional activation by the aryl hydrocarbon receptor. Arch. Biochem. Biophys. 2005, 433, 379–386.
  45. Watzky, M.; Huard, S.; Juricek, L.; Dairou, J.; Chauvet, C.; Coumoul, X.; Letessier, A.; Miotto, B. Hexokinase 2 is a transcriptional target and a positive modulator of AHR signalling. Nucleic Acids Res. 2022, 50, 5545–5564.
  46. Evans, B.R.; Karchner, S.I.; Allan, L.L.; Pollenz, R.S.; Tanguay, R.L.; Jenny, M.J.; Sherr, D.H.; Hahn, M.E. Repression of Aryl Hydrocarbon Receptor (AHR) Signaling by AHR Repressor: Role of DNA Binding and Competition for AHR Nuclear Translocator. Mol. Pharmacol. 2008, 73, 387–398.
  47. Ahmed, S.; Bott, D.; Gomez, A.; Tamblyn, L.; Rasheed, A.; Cho, T.; MacPherson, L.; Sugamori, K.S.; Yang, Y.; Grant, D.M.; et al. Loss of the Mono-ADP-ribosyltransferase, Tiparp, Increases Sensitivity to Dioxin-induced Steatohepatitis and Lethality. J. Biol. Chem. 2015, 290, 16824–16840.
  48. Grimaldi, G.; Rajendra, S.; Matthews, J. The aryl hydrocarbon receptor regulates the expression of TIPARP and its cis long non-coding RNA, TIPARP-AS1. Biochem. Biophys. Res. Commun. 2018, 495, 2356–2362.
  49. Hutin, D.; Tamblyn, L.; Gomez, A.; Grimaldi, G.; Soedling, H.; Cho, T.; Ahmed, S.; Lucas, C.; Kanduri, C.; Grant, D.M.; et al. Hepatocyte-Specific Deletion of TIPARP, a Negative Regulator of the Aryl Hydrocarbon Receptor, Is Sufficient to Increase Sensitivity to Dioxin-Induced Wasting Syndrome. Toxicol. Sci. 2018, 165, 347–360.
  50. Zhang, L.; Cao, J.; Dong, L.; Lin, H. TiPARP forms nuclear condensates to degrade HIF-1α and suppress tumorigenesis. Proc. Natl. Acad. Sci. USA 2020, 117, 13447–13456.
  51. Lu, P.; Cai, X.; Guo, Y.; Xu, M.; Tian, J.; Locker, J.; Xie, W. Constitutive Activation of the Human Aryl Hydrocarbon Receptor in Mice Promotes Hepatocarcinogenesis Independent of Its Coactivator Gadd45b. Toxicol. Sci. 2019, 167, 581–592.
  52. Weinert, B.T.; Narita, T.; Satpathy, S.; Srinivasan, B.; Hansen, B.K.; Schölz, C.; Hamilton, W.B.; Zucconi, B.E.; Wang, W.W.; Liu, W.R.; et al. Time-Resolved Analysis Reveals Rapid Dynamics and Broad Scope of the CBP/p300 Acetylome. Cell 2018, 174, 231–244.e12.
  53. Nakano, N.; Sakata, N.; Katsu, Y.; Nochise, D.; Sato, E.; Takahashi, Y.; Yamaguchi, S.; Haga, Y.; Ikeno, S.; Motizuki, M.; et al. Dissociation of the AhR/ARNT complex by TGF-β/Smad signaling represses CYP1A1 gene expression and inhibits benzepyrene-mediated cytotoxicity. J. Biol. Chem. 2020, 295, 9033–9051.
  54. Solaimani, P.; Damoiseaux, R.; Hankinson, O. Genome-Wide RNAi High-Throughput Screen Identifies Proteins Necessary for the AHR-Dependent Induction of CYP1A1 by 2,3,7,8-Tetrachlorodibenzo-p-dioxin. Toxicol. Sci. 2013, 136, 107–119.
  55. Solaimani, P.; Wang, F.; Hankinson, O. SIN3A, Generally Regarded as a Transcriptional Repressor, Is Required for Induction of Gene Transcription by the Aryl Hydrocarbon Receptor. J. Biol. Chem. 2014, 289, 33655–33662.
  56. Bourner, L.A.; Muro, I.; Cooper, A.M.; Choudhury, B.K.; Bailey, A.O.; Russell, W.K.; Khanipov, K.; Golovko, G.; Wright, C.W. AhR promotes phosphorylation of ARNT isoform 1 in human T cell malignancies as a switch for optimal AhR activity. Proc. Natl. Acad. Sci. USA 2022, 119, e2114336119.
  57. Elson, D.J.; Nguyen, B.D.; Wood, R.; Zhang, Y.; Puig-Sanvicens, V.; Kolluri, S.K. The cyclin-dependent kinase inhibitor p27 Kip1 interacts with the aryl hydrocarbon receptor and negatively regulates its transcriptional activity. FEBS Lett. 2022, 596, 2056–2071.
  58. Orlando, S.; Gallastegui, E.; Besson, A.; Abril, G.; Aligué, R.; Pujol, M.J.; Bachs, O. p27Kip1and p21Cip1collaborate in the regulation of transcription by recruiting cyclin–Cdk complexes on the promoters of target genes. Nucleic Acids Res. 2015, 43, 6860–6873.
  59. Pippa, R.; Espinosa, L.; Gundem, G.; García-Escudero, R.; Dominguez, A.; Orlando, S.; Gallastegui, E.; Saiz, C.; Besson, A.; Pujol, M.J.; et al. p27Kip1 represses transcription by direct interaction with p130/E2F4 at the promoters of target genes. Oncogene 2012, 31, 4207–4220.
  60. Puga, A.; Barnes, S.J.; Dalton, T.P.; Chang, C.-Y.; Knudsen, E.S.; Maier, M.A. Aromatic Hydrocarbon Receptor Interaction with the Retinoblastoma Protein Potentiates Repression of E2F-dependent Transcription and Cell Cycle Arrest. J. Biol. Chem. 2000, 275, 2943–2950.
  61. Elferink, C.J.; Ge, N.-L.; Levine, A. Maximal Aryl Hydrocarbon Receptor Activity Depends on an Interaction with the Retinoblastoma Protein. Mol. Pharmacol. 2001, 59, 664–673.
  62. Perearnau, A.; Orlando, S.; Islam, A.B.; Gallastegui, E.; Martínez, J.; Jordan, A.; Bigas, A.; Aligué, R.; Pujol, M.J.; Bachs, O. p27Kip1, PCAF and PAX5 cooperate in the transcriptional regulation of specific target genes. Nucleic Acids Res. 2017, 45, 5086–5099.
  63. Pérez-Luna, M.; Aguasca, M.; Perearnau, A.; Serratosa, J.; Martínez-Balbas, M.; Pujol, M.J.; Bachs, O. PCAF regulates the stability of the transcriptional regulator and cyclin-dependent kinase inhibitor p27Kip1. Nucleic Acids Res. 2012, 40, 6520–6533.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : ,
View Times: 380
Revisions: 2 times (View History)
Update Date: 19 Apr 2023
1000/1000
Video Production Service