Glucocorticoid and cytokine crosstalk: Feedback, feedforward, and co-regulatory interactions determine repression or resistance

Robert Newton, Suharsh Shah, Mohammed O Altonsy, Antony N Gerber, Robert Newton, Suharsh Shah, Mohammed O Altonsy, Antony N Gerber

Abstract

Inflammatory signals induce feedback and feedforward systems that provide temporal control. Although glucocorticoids can repress inflammatory gene expression, glucocorticoid receptor recruitment increases expression of negative feedback and feedforward regulators, including the phosphatase, DUSP1, the ubiquitin-modifying enzyme, TNFAIP3, or the mRNA-destabilizing protein, ZFP36. Moreover, glucocorticoid receptor cooperativity with factors, including nuclear factor-κB (NF-κB), may enhance regulator expression to promote repression. Conversely, MAPKs, which are inhibited by glucocorticoids, provide feedforward control to limit expression of the transcription factor IRF1, and the chemokine, CXCL10. We propose that modulation of feedback and feedforward control can determine repression or resistance of inflammatory gene expression toglucocorticoid.

Keywords: gene expression; gene regulation; glucocorticoid; glucocorticoid receptor; inflammation.

Conflict of interest statement

The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health

© 2017 by The American Society for Biochemistry and Molecular Biology, Inc.

Figures

Figure 1.
Figure 1.
Differential effects of dexamethasone on IL1B-induced mRNAs. Data are derived from King et al. (11) where full details can be found. A, expression of 39 of the most highly IL1B-induced mRNAs in A549 cells was examined by qPCR following no treatment (NT) or treatment with dexamethasone (1 μm) (Dex), IL1B (1 ng/ml), or IL1B + Dex for 6 h. Data for each mRNA were normalized to GAPDH and are expressed as a percentage of IL1B treated, which is set to 100% (blue), and are presented as a heat map (white = 0%). IL1B-induced mRNAs are ranked according to the effect of dexamethasone with IL6 being the most repressed and CSF3 (G-CSF) being enhanced. The effects of NF-κB inhibition, using adenoviral overexpression of the dominant inhibitor, IκBαΔN, or a control virus (Control), are also expressed as a percentage of IL1B-treated and are shown as a heat map. B, A549 cells were stimulated with IL1B (1 ng/ml) in the absence or presence of the indicated concentrations of dexamethasone prior to harvesting at 6 h for qPCR analysis of the 39 mRNAs in A. Data (n = 6) for each mRNA (indicated by an x) were plotted as a percentage of IL1B treatment. The most potently repressed mRNA (IL1B) along with mRNAs showing intermediate (IFIT3 isoform 2, EFNA1, CFB) and no (IL32) repression or enhancement (CSF3) by dexamethasone are shown (upper panel). The maximal effect (Max. effect) of dexamethasone (i.e. at 1 μm) and the EC50 were calculated for each mRNA that showed significant repression. These are plotted (lower panel) to show the correlation between potency and repression by dexamethasone. Error bars indicate ± S.E.
Figure 2.
Figure 2.
Regulatory loops controlling inflammatory gene expression.A, schematic showing activation of MAPK pathways and NF-κB by IL1B or TNF leading to the expression of inflammatory genes. MAPKs induce the expression of the phosphatase, DUSP1, which provides feedback control to switch off MAPK activity. NF-κB binds κB sites in promoters of target genes. This activates transcription of NFKBIA, TNFAIP3, and IRAK3 to increase their expression and leads to feedback inhibition of NF-κB or IL1B/TNF signaling. Expression of DUSP1, NFKBIA, TNFAIP3, and IRAK3 can also be enhanced by glucocorticoids (GC). B, type I coherent and incoherent feedforward loops are depicted. In the type I coherent feedforward loop (panel i), X positively regulates Y, and Z is positively regulated by both X and Y. In the type I incoherent feedforward loop (panel ii), X positively regulates both Y and Z, but Y negatively regulates Z. C, schematic showing how feedback and feedforward regulation may interplay to regulate AU-rich element (ARE)-containing inflammatory mRNAs. Panel i, pro-inflammatory stimuli, here IL1B, activate MAPK pathways, leading to the expression of ARE-containing mRNAs, such as TNF. MAPK activation not only also induces expression of the feedback regulator, DUSP1, but also promotes expression of ZFP36. ZFP36 is a feedforward regulator that leads to mRNA destabilization of ARE-containing mRNAs, such as TNF. Thus, the MAPK-dependent induction of ZFP36 leads to repression of ARE-containing mRNAs and constitutes a classic type I incoherent feedforward loop. Note that expression of ARE-containing mRNAs and ZFP36 is also likely to involve NF-κB, and this is not depicted. Panel ii, Following loss, or silencing, of DUSP1, MAPK activity is enhanced and leads to increased expression of downstream genes. However, expression of ZFP36 is also enhanced, and this acts to reduce expression of ARE-containing mRNAs, such as TNF. Panel iii, ZFP36 expression is up-regulated by glucocorticoids alone, but ZFP36 expression induced by the inflammatory stimulus is reduced by glucocorticoid, in part due to reduced MAPK activity following the induction of DUSP1. Although these effects may combine to promote expression of the hypo-phosphorylated and more active, mRNA-destabilizing form of ZFP36, silencing of both DUSP1 and ZFP36 showed little effect on the repression of TNF by glucocorticoid. Additional, glucocorticoid-induced effector processes are therefore likely to play additional repressive roles.
Figure 3.
Figure 3.
GR and NF-κB (RELA) recruitment to inflammatory gene loci. Data from a ChIP-seq analysis by Kadiyala et al. (15) are shown. BEAS-2B cells were treated for 1 h with Dex (1 μm), TNF (20 ng/ml), or Dex plus TNF for 1 h prior to ChIP-seq analysis. Occupancy of GR and RELA at genomic loci in the vicinity of six genes is shown. A, inflammatory feedback control genes, where GR and RELA may cooperate to enhance or maintain the expression of TNFAIP3 (panel i); NFKBIA (panel ii); and IRAK3 (panel iii). In each case, GR and RELA are recruited to the gene loci following dexamethasone or TNF treatment, respectively. In the context of dexamethasone plus TNF, both GR and RELA are both recruited to at least one DNA region in common (red arrow). Although overall GR occupancy at each gene locus was largely unaffected by TNF, site-specific differences are apparent. Conversely, co-treatment differentially affected RELA occupancy, which was increased at an intronic region for TNFAIP3, slightly decreased on NFKBIA, and markedly increased at the IRAK3 promoter. B, GR and RELA co-recruitment to the SOD2 (panel i), ZCH12A (panel ii), and IL32 (panel iii) gene loci. SOD2 and IL32 are induced by inflammatory stimuli, and here TNF induces RELA binding to each gene locus. In the additional presence of dexamethasone, RELA binding is slightly reduced (red arrow). However, although GR occupancy at this same region was not readily apparent with dexamethasone alone, with TNF plus dexamethasone, GR recruitment is induced (red arrow). GR occupancy at regions that either did not show GR binding, or only showed weak GR binding, in the presence of dexamethasone alone were also enhanced for both SOD2 and IL32 with dexamethasone plus TNF (black arrows). These regions did not show material RELA occupancy. With ZC3H12A, RELA occupancy was induced to multiple intronic regions by TNF. Although dexamethasone reduced RELA occupancy, GR was recruited to these same regions with TNF plus dexamethasone (red arrows). Binding of RELA and GR at a 5′ region was markedly enhanced by TNF plus dexamethasone (black arrows), whereas neither TNF nor dexamethasone alone showed any marked effect on occupancy.
Figure 4.
Figure 4.
Loss of feedforward control may promote glucocorticoid resistance.A, schematic showing the regulation of IRF1, as well as IRF1-dependent gene expression. Pro-inflammatory stimuli (IL1B) induce NF-κB activity, leading to the transcriptional activation of IRF1. IRF1 expression is rapidly induced to promote expression of downstream IRF1-dependent genes. Many IRF1-depedent genes, for example CXCL10, are also directly regulated by NF-κB to constitute a type I coherent feedforward loop. Pro-inflammatory stimuli, such as IL1B, promote activation of MAPK pathways. Acting via multiple mechanisms, MAPKs promote the switching off and/or loss of IRF1 expression. This terminates IRF1 expression and prevents continued expression of IRF1-dependent genes. In the presence of glucocorticoid (GC), DUSP1 expression is enhanced. By reducing MAPK activity, glucocorticoids reduce incoherent feedforward control of IRF1, and this promotes IRF1 expression. This effect also occurs following MAPK inhibition. Maintenance of IRF1 expression helps CXCL10 to escape the otherwise repressive effects of the glucocorticoid. However, many other IRF1-dependent genes (for examples, see CMPK2, MX1, IFIT1, and others on Fig. 1A) show significant repression by glucocorticoids. Therefore, the existence of additional mechanisms of repression must be invoked. B, data are modified from Shah et al. (77) where full details can be found. A549 cells were either not treated, or treated with IL1B (1 ng/ml), for the times indicated prior to Western blotting and qPCR analysis of IRF1 and GAPDH (upper and middle panels) or qPCR of established IRF1-dependent genes. Spike kinetics for IFR1 mRNA and protein, as well as late-phase kinetics for IRF1-dependent genes, is shown.

References

    1. Oakley R. H., and Cidlowski J. A. (2013) The biology of the glucocorticoid receptor: new signaling mechanisms in health and disease. J. Allergy Clin. Immunol. 132, 1033–1044
    1. Barnes P. J. (2011) Glucocorticosteroids: current and future directions. Br. J. Pharmacol. 163, 29–43
    1. Barnes P. J. (2013) Corticosteroid resistance in patients with asthma and chronic obstructive pulmonary disease. J. Allergy Clin. Immunol. 131, 636–645
    1. Keenan C. R., Salem S., Fietz E. R., Gualano R. C., and Stewart A. G. (2012) Glucocorticoid-resistant asthma and novel anti-inflammatory drugs. Drug Discov. Today 17, 1031–1038
    1. Ammit A. J. (2013) Glucocorticoid insensitivity as a source of drug targets for respiratory disease. Curr. Opin. Pharmacol. 13, 370–376
    1. Sukkar M. B., Issa R., Xie S., Oltmanns U., Newton R., and Chung K. F. (2004) Fractalkine/CX3CL1 production by human airway smooth muscle cells: induction by IFN-γ and TNF-α and regulation by TGF-β and corticosteroids. Am. J. Physiol. Lung Cell Mol. Physiol. 287, L1230–L1240
    1. Zhang N., Truong-Tran Q. A., Tancowny B., Harris K. E., and Schleimer R. P. (2007) Glucocorticoids enhance or spare innate immunity: effects in airway epithelium are mediated by CCAAT/enhancer binding proteins. J. Immunol. 179, 578–589
    1. Langlais D., Couture C., Balsalobre A., and Drouin J. (2008) Regulatory network analyses reveal genome-wide potentiation of LIF signaling by glucocorticoids and define an innate cell defense response. PLoS. Genet. 4, e1000224.
    1. Busillo J. M., Azzam K. M., and Cidlowski J. A. (2011) Glucocorticoids sensitize the innate immune system through regulation of the NLRP3 inflammasome. J. Biol. Chem. 286, 38703–38713
    1. Lannan E. A., Galliher-Beckley A. J., Scoltock A. B., and Cidlowski J. A. (2012) Proinflammatory actions of glucocorticoids: glucocorticoids and TNFα coregulate gene expression in vitro and in vivo. Endocrinology 153, 3701–3712
    1. King E. M., Chivers J. E., Rider C. F., Minnich A., Giembycz M. A., and Newton R. (2013) Glucocorticoid repression of inflammatory gene expression shows differential responsiveness by transactivation- and transrepression-dependent mechanisms. PLoS ONE 8, e53936.
    1. Busillo J. M., and Cidlowski J. A. (2013) The five Rs of glucocorticoid action during inflammation: ready, reinforce, repress, resolve, and restore. Trends Endocrinol. Metab. 24, 109–119
    1. Rao N. A., McCalman M. T., Moulos P., Francoijs K. J., Chatziioannou A., Kolisis F. N., Alexis M. N., Mitsiou D. J., and Stunnenberg H. G. (2011) Coactivation of GR and NFKB alters the repertoire of their binding sites and target genes. Genome Res. 21, 1404–1416
    1. van de Garde M. D., Martinez F. O., Melgert B. N., Hylkema M. N., Jonkers R. E., and Hamann J. (2014) Chronic exposure to glucocorticoids shapes gene expression and modulates innate and adaptive activation pathways in macrophages with distinct changes in leukocyte attraction. J. Immunol. 192, 1196–1208
    1. Kadiyala V., Sasse S. K., Altonsy M. O., Berman R., Chu H. W., Phang T. L., and Gerber A. N. (2016) Cistrome-based cooperation between airway epithelial glucocorticoid receptor and NF-κB orchestrates anti-inflammatory effects. J. Biol. Chem. 291, 12673–12687
    1. Leigh R., Mostafa M. M., King E. M., Rider C. F., Shah S., Dumonceaux C., Traves S. L., McWhae A., Kolisnik T., Kooi C., Slater D. M., Kelly M. M., Bieda M., Miller-Larsson A., and Newton R. (2016) An inhaled dose of budesonide induces genes involved in transcription and signaling in the human airways: enhancement of anti- and proinflammatory effector genes. Pharmacol. Res. Perspect. 4, e00243.
    1. Newton R., and Holden N. S. (2007) Separating transrepression and transactivation: a distressing divorce for the glucocorticoid receptor? Mol. Pharmacol. 72, 799–809
    1. Petta I., Dejager L., Ballegeer M., Lievens S., Tavernier J., De Bosscher K., and Libert C. (2016) The interactome of the glucocorticoid receptor and its influence on the actions of glucocorticoids in combatting inflammatory and infectious diseases. Microbiol. Mol. Biol. Rev. 80, 495–522
    1. Clark A. R., and Belvisi M. G. (2012) Maps and legends: the quest for dissociated ligands of the glucocorticoid receptor. Pharmacol. Ther. 134, 54–67
    1. Ito K., Yamamura S., Essilfie-Quaye S., Cosio B., Ito M., Barnes P. J., and Adcock I. M. (2006) Histone deacetylase 2-mediated deacetylation of the glucocorticoid receptor enables NF-κB suppression. J. Exp. Med. 203, 7–13
    1. Hua G., Ganti K. P., and Chambon P. (2016) Glucocorticoid-induced tethered transrepression requires SUMOylation of GR and formation of a SUMO-SMRT/NCoR1-HDAC3 repressing complex. Proc. Natl. Acad. Sci. U.S.A. 113, E635–E643
    1. Surjit M., Ganti K. P., Mukherji A., Ye T., Hua G., Metzger D., Li M., and Chambon P. (2011) Widespread negative response elements mediate direct repression by agonist-liganded glucocorticoid receptor. Cell 145, 224–241
    1. Hudson W. H., Youn C., and Ortlund E. A. (2013) The structural basis of direct glucocorticoid-mediated transrepression. Nat. Struct. Mol. Biol. 20, 53–58
    1. Hua G., Paulen L., and Chambon P. (2016) GR SUMOylation and formation of an SUMO-SMRT/NCoR1-HDAC3 repressing complex is mandatory for GC-induced IR nGRE-mediated transrepression. Proc. Natl. Acad. Sci. U.S.A. 113, E626–E634
    1. So A. Y., Chaivorapol C., Bolton E. C., Li H., and Yamamoto K. R. (2007) Determinants of cell- and gene-specific transcriptional regulation by the glucocorticoid receptor. PLoS. Genet. 3, e94.
    1. Biddie S. C., John S., Sabo P. J., Thurman R. E., Johnson T. A., Schiltz R. L., Miranda T. B., Sung M. H., Trump S., Lightman S. L., Vinson C., Stamatoyannopoulos J. A., and Hager G. L. (2011) Transcription factor AP1 potentiates chromatin accessibility and glucocorticoid receptor binding. Mol. Cell 43, 145–155
    1. Grøntved L., John S., Baek S., Liu Y., Buckley J. R., Vinson C., Aguilera G., and Hager G. L. (2013) C/EBP maintains chromatin accessibility in liver and facilitates glucocorticoid receptor recruitment to steroid response elements. EMBO J. 32, 1568–1583
    1. Langlais D., Couture C., Balsalobre A., and Drouin J. (2012) The Stat3/GR interaction code: predictive value of direct/indirect DNA recruitment for transcription outcome. Mol. Cell 47, 38–49
    1. Arthur J. S., and Ley S. C. (2013) Mitogen-activated protein kinases in innate immunity. Nat. Rev. Immunol. 13, 679–692
    1. Clark A. R., Martins J. R., and Tchen C. R. (2008) Role of dual specificity phosphatases in biological responses to glucocorticoids. J. Biol. Chem. 283, 25765–25769
    1. Shah S., King E. M., Chandrasekhar A., and Newton R. (2014) Roles for the mitogen-activated protein kinase (MAPK) phosphatase, DUSP1, in feedback control of inflammatory gene expression and repression by dexamethasone. J. Biol. Chem. 289, 13667–13679
    1. Diefenbacher M., Sekula S., Heilbock C., Maier J. V., Litfin M., van Dam H., Castellazzi M., Herrlich P., and Kassel O. (2008) Restriction to Fos family members of Trip6-dependent coactivation and glucocorticoid receptor-dependent trans-repression of activator protein-1. Mol. Endocrinol. 22, 1767–1780
    1. King E. M., Holden N. S., Gong W., Rider C. F., and Newton R. (2009) Inhibition of NF-κB-dependent transcription by MKP-1: transcriptional repression by glucocorticoids occurring via p38 MAPK. J. Biol. Chem. 284, 26803–26815
    1. Bladh L. G., Johansson-Haque K., Rafter I., Nilsson S., and Okret S. (2009) Inhibition of extracellular signal-regulated kinase (ERK) signaling participates in repression of nuclear factor (NF)-κB activity by glucocorticoids. Biochim. Biophys. Acta 1793, 439–446
    1. Issa R., Xie S., Khorasani N., Sukkar M., Adcock I. M., Lee K. Y., and Chung K. F. (2007) Corticosteroid inhibition of growth-related oncogene protein-α via mitogen-activated kinase phosphatase-1 in airway smooth muscle cells. J. Immunol. 178, 7366–7375
    1. Abraham S. M., Lawrence T., Kleiman A., Warden P., Medghalchi M., Tuckermann J., Saklatvala J., and Clark A. R. (2006) Antiinflammatory effects of dexamethasone are partly dependent on induction of dual specificity phosphatase 1. J. Exp. Med. 203, 1883–1889
    1. Maier J. V., Brema S., Tuckermann J., Herzer U., Klein M., Stassen M., Moorthy A., and Cato A. C. (2007) Dual specificity phosphatase 1 knockout mice show enhanced susceptibility to anaphylaxis but are sensitive to glucocorticoids. Mol. Endocrinol. 21, 2663–2671
    1. Lasa M., Abraham S. M., Boucheron C., Saklatvala J., and Clark A. R. (2002) Dexamethasone causes sustained expression of mitogen-activated protein kinase (MAPK) phosphatase 1 and phosphatase-mediated inhibition of MAPK p38. Mol. Cell Biol. 22, 7802–7811
    1. Imasato A., Desbois-Mouthon C., Han J., Kai H., Cato A. C., Akira S., and Li J. D. (2002) Inhibition of p38 MAPK by glucocorticoids via induction of MAPK phosphatase-1 enhances nontypeable Haemophilus influenzae-induced expression of Toll-like receptor 2. J. Biol. Chem. 277, 47444–47450
    1. Newton R., King E. M., Gong W., Rider C. F., Staples K. J., Holden N. S., and Bergmann M. W. (2010) Glucocorticoids inhibit IL-1β-induced GM-CSF expression at multiple levels: roles for the ERK pathway and repression by MKP-1. Biochem. J. 427, 113–124
    1. Johansson-Haque K., Palanichamy E., and Okret S. (2008) Stimulation of MAPK-phosphatase 1 gene expression by glucocorticoids occurs through a tethering mechanism involving C/EBP. J. Mol. Endocrinol. 41, 239–249
    1. Tchen C. R., Martins J. R., Paktiawal N., Perelli R., Saklatvala J., and Clark A. R. (2010) Glucocorticoid regulation of mouse and human dual specificity phosphatase 1 (DUSP1) genes: unusual cis-acting elements and unexpected evolutionary divergence. J. Biol. Chem. 285, 2642–2652
    1. Vandevyver S., Dejager L., Van Bogaert T., Kleyman A., Liu Y., Tuckermann J., and Libert C. (2012) Glucocorticoid receptor dimerization induces MKP1 to protect against TNF-induced inflammation. J. Clin. Invest. 122, 2130–2140
    1. Shipp L. E., Lee J. V., Yu C. Y., Pufall M., Zhang P., Scott D. K., and Wang J. C. (2010) Transcriptional regulation of human dual specificity protein phosphatase 1 (DUSP1) gene by glucocorticoids. PLoS. One 5, e13754.
    1. Jubb A. W., Young R. S., Hume D. A., and Bickmore W. A. (2016) Enhancer turnover is associated with a divergent transcriptional response to glucocorticoid in mouse and human macrophages. J. Immunol. 196, 813–822
    1. Wertz I. E., O'Rourke K. M., Zhou H., Eby M., Aravind L., Seshagiri S., Wu P., Wiesmann C., Baker R., Boone D. L., Ma A., Koonin E. V., and Dixit V. M. (2004) De-ubiquitination and ubiquitin ligase domains of A20 downregulate NF-κB signalling. Nature 430, 694–699
    1. Tokunaga F. (2013) Linear ubiquitination-mediated NF-κB regulation and its related disorders. J. Biochem. 154, 313–323
    1. Altonsy M. O., Sasse S. K., Phang T. L., and Gerber A. N. (2014) Context-dependent cooperation between nuclear factor κB (NF-κB) and the glucocorticoid receptor at a TNFAIP3 intronic enhancer: a mechanism to maintain negative feedback control of inflammation. J. Biol. Chem. 289, 8231–8239
    1. Sasse S. K., Altonsy M. O., Kadiyala V., Cao G., Panettieri R. A. Jr., and Gerber A. N. (2016) Glucocorticoid and TNF signaling converge at A20 (TNFAIP3) to repress airway smooth muscle cytokine expression. Am. J. Physiol. Lung Cell Mol. Physiol. 311, L421–L432
    1. Hofmann T. G., and Schmitz M. L. (2002) The promoter context determines mutual repression or synergism between NF-κB and the glucocorticoid receptor. Biol. Chem. 383, 1947–1951
    1. Le Bail O., Schmidt-Ullrich R., and Israël A. (1993) Promoter analysis of the gene encoding the IκB-α/MAD3 inhibitor of NF-κB: positive regulation by members of the rel/NF-κB family. EMBO J. 12, 5043–5049
    1. Newton R. (2014) Anti-inflammatory glucocorticoids: changing concepts. Eur. J. Pharmacol. 724, 231–236
    1. Reddy T. E., Pauli F., Sprouse R. O., Neff N. F., Newberry K. M., Garabedian M. J., and Myers R. M. (2009) Genomic determination of the glucocorticoid response reveals unexpected mechanisms of gene regulation. Genome Res. 19, 2163–2171
    1. Newton R., Hart L. A., Stevens D. A., Bergmann M., Donnelly L. E., Adcock I. M., and Barnes P. J. (1998) Effect of dexamethasone on interleukin-1β-(IL-1β)-induced nuclear factor-κB (NF-κB) and κB-dependent transcription in epithelial cells. Eur. J. Biochem. 254, 81–89
    1. Hubbard L. L., and Moore B. B. (2010) IRAK-M regulation and function in host defense and immune homeostasis. Infect. Dis. Rep. 2, e9.
    1. Miyata M., Lee J. Y., Susuki-Miyata S., Wang W. Y., Xu H., Kai H., Kobayashi K. S., Flavell R. A., and Li J. D. (2015) Glucocorticoids suppress inflammation via the upregulation of negative regulator IRAK-M. Nat. Commun. 6, 6062.
    1. Alon U. (2007) Network motifs: theory and experimental approaches. Nat. Rev. Genet. 8, 450–461
    1. Sasse S. K., and Gerber A. N. (2015) Feed-forward transcriptional programming by nuclear receptors: regulatory principles and therapeutic implications. Pharmacol. Ther. 145, 85–91
    1. Anderson P. (2008) Post-transcriptional control of cytokine production. Nat. Immunol. 9, 353–359
    1. Clark A. R., and Dean J. L. (2016) The control of inflammation via the phosphorylation and dephosphorylation of tristetraprolin: a tale of two phosphatases. Biochem. Soc. Trans. 44, 1321–1337
    1. King E. M., Kaur M., Gong W., Rider C. F., Holden N. S., and Newton R. (2009) Regulation of tristetraprolin expression by interleukin-1β and dexamethasone in human pulmonary epithelial cells: roles for nuclear factor-κB and p38 mitogen-activated protein kinase. J. Pharmacol. Exp. Ther. 330, 575–585
    1. Prabhala P., Bunge K., Rahman M. M., Ge Q., Clark A. R., and Ammit A. J. (2015) Temporal regulation of cytokine mRNA expression by tristetraprolin: dynamic control by p38 MAPK and MKP-1. Am. J. Physiol. Lung Cell Mol. Physiol. 308, L973–L980
    1. Shah S., Mostafa M. M., McWhae A., Traves S. L., and Newton R. (2016) Negative feed-forward control of tumor necrosis factor (TNF) by tristetraprolin (ZFP36) is limited by the mitogen-activated protein kinase phosphatase, dual-specificity phosphatase 1 (DUSP1): implications for regulation by glucocorticoids. J. Biol. Chem. 291, 110–125
    1. Huotari N., Hömmö T., Taimi V., Nieminen R., Moilanen E., and Korhonen R. (2012) Regulation of tristetraprolin expression by mitogen-activated protein kinase phosphatase-1. APMIS 120, 988–999
    1. Ross E. A., Smallie T., Ding Q., O'Neil J. D., Cunliffe H. E., Tang T., Rosner D. R., Klevernic I., Morrice N. A., Monaco C., Cunningham A. F., Buckley C. D., Saklatvala J., Dean J. L., and Clark A. R. (2015) Dominant suppression of inflammation via targeted mutation of the mRNA destabilizing protein tristetraprolin. J. Immunol. 195, 265–276
    1. Smallie T., Ross E. A., Ammit A. J., Cunliffe H. E., Tang T., Rosner D. R., Ridley M. L., Buckley C. D., Saklatvala J., Dean J. L., and Clark A. R. (2015) Dual-specificity phosphatase 1 and tristetraprolin cooperate to regulate macrophage responses to lipopolysaccharide. J. Immunol. 195, 277–288
    1. Kratochvill F., Machacek C., Vogl C., Ebner F., Sedlyarov V., Gruber A. R., Hartweger H., Vielnascher R., Karaghiosoff M., Rülicke T., Müller M., Hofacker I., Lang R., and Kovarik P. (2011) Tristetraprolin-driven regulatory circuit controls quality and timing of mRNA decay in inflammation. Mol. Syst. Biol. 7, 560.
    1. Sun L., Stoecklin G., Van Way S., Hinkovska-Galcheva V., Guo R. F., Anderson P., and Shanley T. P. (2007) Tristetraprolin (TTP)-14-3-3 complex formation protects TTP from dephosphorylation by protein phosphatase 2a and stabilizes tumor necrosis factor-α mRNA. J. Biol. Chem. 282, 3766–3777
    1. Rahman M. M., Rumzhum N. N., Morris J. C., Clark A. R., Verrills N. M., and Ammit A. J. (2015) Basal protein phosphatase 2A activity restrains cytokine expression: role for MAPKs and tristetraprolin. Sci. Rep. 5, 10063.
    1. Jalonen U., Lahti A., Korhonen R., Kankaanranta H., and Moilanen E. (2005) Inhibition of tristetraprolin expression by dexamethasone in activated macrophages. Biochem. Pharmacol. 69, 733–740
    1. Prabhala P., Bunge K., Ge Q., and Ammit A. J. (2016) Corticosteroid-induced MKP-1 represses pro-inflammatory cytokine secretion by enhancing activity of tristetraprolin (TTP) in ASM cells. J. Cell Physiol. 231, 2153–2158
    1. Chivers J. E., Gong W., King E. M., Seybold J., Mak J. C., Donnelly L. E., Holden N. S., and Newton R. (2006) Analysis of the dissociated steroid, RU24858, does not exclude a role for inducible genes in the anti-inflammatory actions of glucocorticoids. Mol. Pharmacol. 70, 2084–2095
    1. Smoak K., and Cidlowski J. A. (2006) Glucocorticoids regulate tristetraprolin synthesis and posttranscriptionally regulate tumor necrosis factor α inflammatory signaling. Mol. Cell Biol. 26, 9126–9135
    1. Kaur M., Chivers J. E., Giembycz M. A., and Newton R. (2008) Long-acting β2-adrenoceptor agonists synergistically enhance glucocorticoid-dependent transcription in human airway epithelial and smooth muscle cells. Mol. Pharmacol. 73, 203–214
    1. So A. Y., Cooper S. B., Feldman B. J., Manuchehri M., and Yamamoto K. R. (2008) Conservation analysis predicts in vivo occupancy of glucocorticoid receptor-binding sequences at glucocorticoid-induced genes. Proc. Natl. Acad. Sci. U.S.A. 105, 5745–5749
    1. Tiedje C., Diaz-Muñoz M. D., Trulley P., Ahlfors H., Laaß K., Blackshear P. J., Turner M., and Gaestel M. (2016) The RNA-binding protein TTP is a global post-transcriptional regulator of feedback control in inflammation. Nucleic Acids Res. 44, 7418–7440
    1. Shah S., King E. M., Mostafa M. M., Altonsy M. O., and Newton R. (2016) DUSP1 maintains IRF1 and leads to increased expression of IRF1-dependent genes: a mechanism promoting glucocorticoid insensitivity. J. Biol. Chem. 291, 21802–21816
    1. Hammer M., Mages J., Dietrich H., Servatius A., Howells N., Cato A. C., and Lang R. (2006) Dual specificity phosphatase 1 (DUSP1) regulates a subset of LPS-induced genes and protects mice from lethal endotoxin shock. J. Exp. Med. 203, 15–20
    1. Zaheer R. S., Koetzler R., Holden N. S., Wiehler S., and Proud D. (2009) Selective transcriptional down-regulation of human rhinovirus-induced production of CXCL10 from airway epithelial cells via the MEK1 pathway. J. Immunol. 182, 4854–4864
    1. Korhonen R., Huotari N., Hömmö T., Leppänen T., and Moilanen E. (2012) The expression of interleukin-12 is increased by MAP kinase phosphatase-1 through a mechanism related to interferon regulatory factor 1. Mol. Immunol. 51, 219–226
    1. AbuSara N., Razavi S., Derwish L., Komatsu Y., Licursi M., and Hirasawa K. (2015) Restoration of IRF1-dependent anticancer effects by MEK inhibition in human cancer cells. Cancer Lett. 357, 575–581
    1. Taniguchi T., Ogasawara K., Takaoka A., and Tanaka N. (2001) IRF family of transcription factors as regulators of host defense. Annu. Rev. Immunol. 19, 623–655
    1. Soye K. J., Trottier C., Richardson C. D., Ward B. J., and Miller W. H. Jr. (2011) RIG-I is required for the inhibition of measles virus by retinoids. PLoS. One 6, e22323.
    1. Zaheer R. S., and Proud D. (2010) Human rhinovirus-induced epithelial production of CXCL10 is dependent upon IFN regulatory factor-1. Am. J. Respir. Cell Mol. Biol. 43, 413–421
    1. Leigh R., and Proud D. (2015) Virus-induced modulation of lower airway diseases: pathogenesis and pharmacologic approaches to treatment. Pharmacol. Ther. 148, 185–198
    1. Nair S., Michaelsen-Preusse K., Finsterbusch K., Stegemann-Koniszewski S., Bruder D., Grashoff M., Korte M., Köster M., Kalinke U., Hauser H., and Kröger A. (2014) Interferon regulatory factor-1 protects from fatal neurotropic infection with vesicular stomatitis virus by specific inhibition of viral replication in neurons. PLoS. Pathog. 10, e1003999.
    1. Remoli A. L., Marsili G., Perrotti E., Acchioni C., Sgarbanti M., Borsetti A., Hiscott J., and Battistini A. (2016) HIV-1 Tat recruits HDM2 E3 ligase to target IRF-1 for ubiquitination and proteasomal degradation. MBio. 7, e01528–16
    1. Lin R., and Hiscott J. (1999) A role for casein kinase II phosphorylation in the regulation of IRF-1 transcriptional activity. Mol. Cell. Biochem. 191, 169–180
    1. Nakagawa K., and Yokosawa H. (2000) Degradation of transcription factor IRF-1 by the ubiquitin-proteasome pathway. The C-terminal region governs the protein stability. Eur. J. Biochem. 267, 1680–1686
    1. Sauty A., Dziejman M., Taha R. A., Iarossi A. S., Neote K., Garcia-Zepeda E. A., Hamid Q., and Luster A. D. (1999) The T cell-specific CXC chemokines IP-10, Mig, and I-TAC are expressed by activated human bronchial epithelial cells. J. Immunol. 162, 3549–3558
    1. Wark P. A., Bucchieri F., Johnston S. L., Gibson P. G., Hamilton L., Mimica J., Zummo G., Holgate S. T., Attia J., Thakkinstian A., and Davies D. E. (2007) IFN-γ-induced protein 10 is a novel biomarker of rhinovirus-induced asthma exacerbations. J. Allergy Clin. Immunol. 120, 586–593
    1. Medoff B. D., Sauty A., Tager A. M., Maclean J. A., Smith R. N., Mathew A., Dufour J. H., and Luster A. D. (2002) IFN-γ-inducible protein 10 (CXCL10) contributes to airway hyperreactivity and airway inflammation in a mouse model of asthma. J. Immunol. 168, 5278–5286
    1. Tliba O., Damera G., Banerjee A., Gu S., Baidouri H., Keslacy S., and Amrani Y. (2008) Cytokines induce an early steroid resistance in airway smooth muscle cells: novel role of interferon regulatory factor-1. Am. J. Respir. Cell Mol. Biol. 38, 463–472
    1. Bhandare R., Damera G., Banerjee A., Flammer J. R., Keslacy S., Rogatsky I., Panettieri R. A., Amrani Y., and Tliba O. (2010) Glucocorticoid receptor interacting protein-1 restores glucocorticoid responsiveness in steroid-resistant airway structural cells. Am. J. Respir. Cell Mol. Biol. 42, 9–15
    1. Mylona A., Theillet F. X., Foster C., Cheng T. M., Miralles F., Bates P. A., Selenko P., and Treisman R. (2016) Opposing effects of Elk-1 multisite phosphorylation shape its response to ERK activation. Science 354, 233–237
    1. Laine A., and Ronai Z. (2005) Ubiquitin chains in the ladder of MAPK signaling. Sci. STKE. 2005, re5.
    1. Zhou X., Richon V. M., Wang A. H., Yang X. J., Rifkind R. A., and Marks P. A. (2000) Histone deacetylase 4 associates with extracellular signal-regulated kinases 1 and 2, and its cellular localization is regulated by oncogenic Ras. Proc. Natl. Acad. Sci. U.S.A. 97, 14329–14333
    1. Nguyen L. K., Kolch W., and Kholodenko B. N. (2013) When ubiquitination meets phosphorylation: a systems biology perspective of EGFR/MAPK signalling. Cell Commun. Signal. 11, 52.
    1. Yang S. H., Sharrocks A. D., and Whitmarsh A. J. (2013) MAP kinase signalling cascades and transcriptional regulation. Gene. 513, 1–13

Source: PubMed

3
Prenumerera