An NF-κB-microRNA regulatory network tunes macrophage inflammatory responses

Mati Mann, Arnav Mehta, Jimmy L Zhao, Kevin Lee, Georgi K Marinov, Yvette Garcia-Flores, Li-Fan Lu, Alexander Y Rudensky, David Baltimore, Mati Mann, Arnav Mehta, Jimmy L Zhao, Kevin Lee, Georgi K Marinov, Yvette Garcia-Flores, Li-Fan Lu, Alexander Y Rudensky, David Baltimore

Abstract

The innate inflammatory response must be tightly regulated to ensure effective immune protection. NF-κB is a key mediator of the inflammatory response, and its dysregulation has been associated with immune-related malignancies. Here, we describe a miRNA-based regulatory network that enables precise NF-κB activity in mouse macrophages. Elevated miR-155 expression potentiates NF-κB activity in miR-146a-deficient mice, leading to both an overactive acute inflammatory response and chronic inflammation. Enforced miR-155 expression overrides miR-146a-mediated repression of NF-κB activation, thus emphasizing the dominant function of miR-155 in promoting inflammation. Moreover, miR-155-deficient macrophages exhibit a suboptimal inflammatory response when exposed to low levels of inflammatory stimuli. Importantly, we demonstrate a temporal asymmetry between miR-155 and miR-146a expression during macrophage activation, which creates a combined positive and negative feedback network controlling NF-κB activity. This miRNA-based regulatory network enables a robust yet time-limited inflammatory response essential for functional immunity.MicroRNAs (miR) are important regulators of gene transcription, with miR-155 and miR-146a both implicated in macrophage activation. Here the authors show that NF-κB signalling, miR-155 and miR-146a form a complex network of cross-regulations to control gene transcription in macrophages for modulating inflammatory responses.

Conflict of interest statement

The authors declare no competing financial interests.

Figures

Fig. 1
Fig. 1
miR-155 is required for myeloproliferation and extramedullary haematopoiesis in aged miR-146a−/− mice. a, b Twelve-month-old WT, miR-155−/−, miR-146a−/− and double knock out (DKO) mice were analysed for serum IL-6 levels (a), spleen mRNA levels of IL-1b, IL-6 and TNF (b). cf Twelve-month-old WT, miR-155−/−, miR-146a−/−, DKO and LyzM-Cre miR-146afl/fl mice were analysed for spleen weight (c), relative macrophage percentage in peripheral blood (d) and spleen (e), as well as extramedullary haematopoiesis by HSC quantification in spleen (f). N = 10 (a, b); N = 7 (c, d) per group, from at least two independent experiments and are represented as mean ± SEM. *p < 0.05, **p < 0.01 and ***p < 0.001 using one-way ANOVA
Fig. 2
Fig. 2
miR-155 expression is required for the elevated acute inflammatory response in miR-146a deficient mice. ac Eight to ten-week-old WT, miR-155−/−, miR-146a−/− and double knock out (DKO) mice were infected with Listeria monocytogenes. Seventy-two hours after infection, mice were analysed for colony formation units (CFUs) in spleen (a) and liver (b), as well as serum IL-6 levels (c). Eight to ten-week-old WT, miR-155−/−, miR-146a−/− and DKO mice received 1 mg/kg LPS every 24 h for 3 days via IP injections. Serum IL-6 levels (d) as well as peripheral blood macrophage activation was quantified 4 h after each injection by of CD11b+ MHC-II high (e) and CD11b+ CD80+ percentage (f). Arrows indicate time of LPS injections. gh Time course expression profile of miR-155 of WT and miR-146a−/− (g) and miR-146a of WT and miR-155−/− BMMs (h) after LPS stimulation (100 ng/mL) were quantified by qPCR. N > 7 (ad), or N > 9 (eh) per group from two independent experiments and are represented as mean ± SEM. *p < 0.05, **p < 0.01 and ***p < 0.001 using one-way ANOVA (ad) or two-way ANOVA (dh)
Fig. 3
Fig. 3
MiR-155 overexpression overrides miR-146a repression of NF-κB, leading to NF-κB activation and myeloproliferation. a, b Mice reconstituted with bone marrow transduced with GFP and control (MG), miR-155, miR-146a or both miR-155 and miR-146a (dmiR) were analysed 4 months after reconstitution for (a) CD11b+ peripheral blood macrophages and (b) spleen weight. ce Mice reconstituted with bone marrow from NF-κB reporter mice, transduced with control (MG), miR-155, miR-146a or dmiR were analysed 3 months after reconstitution. (c) Basal NF-κB activity of CD11b+ peripheral blood macrophages. (d) Mice received 1 mg/kg LPS every 24 h for 3 days via IP injections. Peripheral blood macrophage NF-κB activation was quantified 4 h after each injection by GFP expression, as well as serum IL-6 levels (e). Dashed line represent ELISA detection sensitivity. N = 7 per group from three (ac) or two (d, e) independent experiments and are represented as percentage (a) or mean ± SEM. *p < 0.05, **p < 0.01, using one-way ANOVA (b, c) or two-way ANOVA (de)
Fig. 4
Fig. 4
Molecular characterization of miR-146a and miR-155 KO BMMs reveal SHIP1 and SOCS1 as the major miR-155 targets. mRNA from BMMs of WT, miR-155−/−, miR-146a−/− and double knock out (DKO) mice before and 8 h after LPS stimulation was subjected to RNA sequencing. a Heat map, indicating the number of differentially expressed genes between samples of all strains before (top) and 8 h after LPS stimulation (bottom). b Enriched functional annotations for genes upregulated in stimulated miR-146a−/− BMMs compared to WT 8 h after LPS stimulation. c qPCR quantification of potential miR-146a and miR-155 targets in BMMs 24 h after LPS stimulation. d Protein from BMMs of WT, miR-155−/−, miR-146a−/− and dKO mice was subjected to western blot analysis 24 h after LPS stimulation. eh Mice were reconstituted with bone marrow from WT, miR-155−/−, miR-146a−/− or dKO mice transduced with GFP expressing control (MG), SHIP-1 or SOCS-1 shRNA vectors. e Reconstitution competitiveness of SOCS1 and SHIP1 attenuated WT bone marrow cells compared to control as measured by GFP expressing CD45+ cells. f Ki67 and Hochst proliferative state of CD45+ WT cells transduced with SOCS1 shRNA compared to control. g CD11b+ GFP+ peripheral blood macrophage percentage of WT cells transduced with control, SHIP1 shRNA or SOCS1 shRNA. h CD11b+ peripheral macrophage percentage comparison between WT, miR-155−/−, miR-146a−/− and dKO mice transduced with SHIP1 shRNA vector. (a, b) N = 10, (c, d) N = 8 from three independent experiments, (eh) N > 7 per group from two independent experiments. c, e, g, h Presented mean ± SEM. *p < 0.05, **p < 0.01, using one-way ANOVA
Fig. 5
Fig. 5
SHIP1 downregulation restores miR-146a−/− phenotype in DKO macrophages. WT, miR-155−/−, miR-146a−/− and double knock out (DKO) bone marrow cells transduced with control (MG) (left) or SHIP1 shRNA (right) were reconstituted into recipient mice. Three months after reconstitution, the mice received 1 mg/kg LPS every 24 h for 3 days via IP injections. Peripheral blood macrophage CD11b+ percentage (a, b), MHC-II high— (c, d) and CD80+ (e, f) were compared for all strains by FACS analysis. Arrows indicate time of LPS injections. N > 7 per group from two independent experiments, represented as mean ± SEM. *p < 0.05, using two-way ANOVA
Fig. 6
Fig. 6
Mir-155 acts as an amplifier in suboptimal stimulation. WT, miR-155−/−, miR-146a−/− and double knock out (DKO) BMMs were stimulated with low levels of LPS (10 ng/mL) (a) or Pam3CSK4 (5 ng/mL) (b). mRNA levels of IL-1β (left) and IL-6 (right) were then quantified using qRT-PCR. c TNF and phospho-p65 (P-p65) protein expression of BMMs from WT (grey), miR-155−/− (blue), miR-146a−/− (red) and dKO (green) 4 h after stimulation with low levels of LPS (10 ng/mL) as measured by FACS using intracellular staining. N = 4 per group from at least two independent experiments, represented as mean ± SEM. a, b *p < 0.05, using two-way ANOVA
Fig. 7
Fig. 7
MiR-155 and miR-146a form a combined positive and negative auto regulatory loop to control precise NF-κB activity during inflammatory stimuli. The dynamic expression of miR-155, miR-146a, and the NF-κB targets IL-1β and IL-6 at different time points after LPS (a) and Pam3CSK4 (b) stimulation. The dynamic expression of IL-6 (c) and IL-1β (d) mRNA in WT BMMs transfected with miR-146a specific anti miR (red) or control (black) 2 h post LPS stimulation (100 ng/mL). e A suggested model for the miR-146a and miR-155 regulatory loop during inflammatory stimulation. a, bN = 4 per group from two independent experiments. *p < 0.05, using one-way ANOVA (c, d)

References

    1. Kawai T, Akira S. TLR signaling. Cell Death Differ. 2006;13:816–825. doi: 10.1038/sj.cdd.4401850.
    1. Newton K, Dixit VM. Signaling in innate immunity and inflammation. Cold Spring Harb. Perspect. Biol. 2012;4:a006049. doi: 10.1101/cshperspect.a006049.
    1. Karin M, Lawrence T, Nizet V. Innate immunity gone awry: linking microbial infections to chronic inflammation and cancer. Cell. 2006;124:823–835. doi: 10.1016/j.cell.2006.02.016.
    1. Coussens LM, Werb Z. Inflammation and cancer. Nature. 2002;420:860–867. doi: 10.1038/nature01322.
    1. Covert MW, Leung TH, Gaston JE, Baltimore D. Achieving stability of lipopolysaccharide-induced NF-kappaB activation. Science. 2005;309:1854–1857. doi: 10.1126/science.1112304.
    1. Liew FY, Xu D, Brint EK, O’Neill LA. Negative regulation of toll-like receptor-mediated immune responses. Nat. Rev. Immunol. 2005;5:446–458. doi: 10.1038/nri1630.
    1. Baltimore D. NF-kappaB is 25. Nat. Immunol. 2011;12:683–685. doi: 10.1038/ni.2072.
    1. Smale ST. Hierarchies of NF-kappaB target-gene regulation. Nat. Immunol. 2011;12:689–694. doi: 10.1038/ni.2070.
    1. Boldin MP, Baltimore D. MicroRNAs, new effectors and regulators of NF-kappaB. Immunol. Rev. 2012;246:205–220. doi: 10.1111/j.1600-065X.2011.01089.x.
    1. Iliopoulos D, Hirsch HA, Struhl K. An epigenetic switch involving NF-kappaB, Lin28, Let-7 MicroRNA, and IL6 links inflammation to cell transformation. Cell. 2009;139:693–706. doi: 10.1016/j.cell.2009.10.014.
    1. Mehta A, Baltimore D. MicroRNAs as regulatory elements in immune system logic. Nat. Rev. Immunol. 2016;16:279–294. doi: 10.1038/nri.2016.40.
    1. Taganov KD, Boldin MP, Chang KJ, Baltimore D. NF-kappaB-dependent induction of microRNA miR-146, an inhibitor targeted to signaling proteins of innate immune responses. Proc. Natl Acad. Sci. USA. 2006;103:12481–12486. doi: 10.1073/pnas.0605298103.
    1. Zhao JL, et al. NF-kappaB dysregulation in microRNA-146a-deficient mice drives the development of myeloid malignancies. Proc. Natl Acad. Sci. USA. 2011;108:9184–9189. doi: 10.1073/pnas.1105398108.
    1. O’Connell RM, et al. MicroRNA-155 promotes autoimmune inflammation by enhancing inflammatory T cell development. Immunity. 2010;33:607–619. doi: 10.1016/j.immuni.2010.09.009.
    1. Thai TH, et al. Regulation of the germinal center response by microRNA-155. Science. 2007;316:604–608. doi: 10.1126/science.1141229.
    1. Gatto G, et al. Epstein-Barr virus latent membrane protein 1 trans-activates miR-155 transcription through the NF-kappaB pathway. Nucleic Acids Res. 2008;36:6608–6619. doi: 10.1093/nar/gkn666.
    1. Boldin MP, et al. miR-146a is a significant brake on autoimmunity, myeloproliferation, and cancer in mice. J. Exp. Med. 2011;208:1189–1201. doi: 10.1084/jem.20101823.
    1. Nahid MA, Pauley KM, Satoh M, Chan EK. miR-146a is critical for endotoxin-induced tolerance: implication in innate immunity. J. Biol. Chem. 2009;284:34590–34599. doi: 10.1074/jbc.M109.056317.
    1. Vigorito E, Kohlhaas S, Lu D, Leyland R. miR-155: an ancient regulator of the immune system. Immunol. Rev. 2013;253:146–157. doi: 10.1111/imr.12057.
    1. Mann M, Barad O, Agami R, Geiger B, Hornstein E. miRNA-based mechanism for the commitment of multipotent progenitors to a single cellular fate. Proc. Natl Acad. Sci. USA. 2010;107:15804–15809. doi: 10.1073/pnas.0915022107.
    1. Rodriguez A, et al. Requirement of bic/microRNA-155 for normal immune function. Science. 2007;316:608–611. doi: 10.1126/science.1139253.
    1. Eis PS, et al. Accumulation of miR-155 and BIC RNA in human B cell lymphomas. Proc. Natl Acad. Sci. USA. 2005;102:3627–3632. doi: 10.1073/pnas.0500613102.
    1. O’Connell RM, et al. Sustained expression of microRNA-155 in hematopoietic stem cells causes a myeloproliferative disorder. J. Exp. Med. 2008;205:585–594. doi: 10.1084/jem.20072108.
    1. Hu R, et al. miR-155 promotes T follicular helper cell accumulation during chronic, low-grade inflammation. Immunity. 2014;41:605–619. doi: 10.1016/j.immuni.2014.09.015.
    1. Huffaker TB, et al. Epistasis between microRNAs 155 and 146a during T cell-mediated antitumor immunity. Cell Rep. 2012;2:1697–1709. doi: 10.1016/j.celrep.2012.10.025.
    1. Mackaness GB. The influence of immunologically committed lymphoid cells on macrophage activity in vivo. J. Exp. Med. 1969;129:973–992. doi: 10.1084/jem.129.5.973.
    1. Magness ST, et al. In vivo pattern of lipopolysaccharide and anti-CD3-induced NF-kappa B activation using a novel gene-targeted enhanced GFP reporter gene mouse. J. Immunol. 2004;173:1561–1570. doi: 10.4049/jimmunol.173.3.1561.
    1. Lewis BP, Burge CB, Bartel DP. Conserved seed pairing, often flanked by adenosines, indicates that thousands of human genes are microRNA targets. Cell. 2005;120:15–20. doi: 10.1016/j.cell.2004.12.035.
    1. Sly LM, Rauh MJ, Kalesnikoff J, Song CH, Krystal G. LPS-induced upregulation of SHIP is essential for endotoxin tolerance. Immunity. 2004;21:227–239. doi: 10.1016/j.immuni.2004.07.010.
    1. Strebovsky J, Walker P, Lang R, Dalpke AH. Suppressor of cytokine signaling 1 (SOCS1) limits NFkappaB signaling by decreasing p65 stability within the cell nucleus. FASEB J. 2011;25:863–874. doi: 10.1096/fj.10-170597.
    1. Lioubin MN, et al. p150Ship, a signal transduction molecule with inositol polyphosphate-5-phosphatase activity. Genes Dev. 1996;10:1084–1095. doi: 10.1101/gad.10.9.1084.
    1. Ojaniemi M, et al. Phosphatidylinositol 3-kinase is involved in Toll-like receptor 4-mediated cytokine expression in mouse macrophages. Eur. J. Immunol. 2003;33:597–605. doi: 10.1002/eji.200323376.
    1. Kalesnikoff J, et al. SHIP negatively regulates IgE + antigen-induced IL-6 production in mast cells by inhibiting NF-kappa B activity. J. Immunol. 2002;168:4737–4746. doi: 10.4049/jimmunol.168.9.4737.
    1. Damen JE, et al. The 145-kDa protein induced to associate with Shc by multiple cytokines is an inositol tetraphosphate and phosphatidylinositol 3,4,5-triphosphate 5-phosphatase. Proc. Natl Acad. Sci. USA. 1996;93:1689–1693. doi: 10.1073/pnas.93.4.1689.
    1. Beraud C, Henzel WJ, Baeuerle PA. Involvement of regulatory and catalytic subunits of phosphoinositide 3-kinase in NF-kappaB activation. Proc. Natl Acad. Sci. USA. 1999;96:429–434. doi: 10.1073/pnas.96.2.429.
    1. Cekic C, et al. MyD88-dependent SHIP1 regulates proinflammatory signaling pathways in dendritic cells after monophosphoryl lipid A stimulation of TLR4. J. Immunol. 2011;186:3858–3865. doi: 10.4049/jimmunol.1001034.
    1. Ozes ON, et al. NF-kappaB activation by tumour necrosis factor requires the Akt serine-threonine kinase. Nature. 1999;401:82–85. doi: 10.1038/43466.
    1. Hoesel B, Schmid JA. The complexity of NF-kappaB signaling in inflammation and cancer. Mol. Cancer. 2013;12:86. doi: 10.1186/1476-4598-12-86.
    1. Dan HC, et al. Akt-dependent regulation of NF-{kappa}B is controlled by mTOR and Raptor in association with IKK. Genes Dev. 2008;22:1490–1500. doi: 10.1101/gad.1662308.
    1. Casey LC, Balk RA, Bone RC. Plasma cytokine and endotoxin levels correlate with survival in patients with the sepsis syndrome. Ann. Intern. Med. 1993;119:771–778. doi: 10.7326/0003-4819-119-8-199310150-00001.
    1. Cagatay T, Turcotte M, Elowitz MB, Garcia-Ojalvo J, Suel GM. Architecture-dependent noise discriminates functionally analogous differentiation circuits. Cell. 2009;139:512–522. doi: 10.1016/j.cell.2009.07.046.
    1. Locke JC, Young JW, Fontes M, Hernandez Jimenez MJ, Elowitz MB. Stochastic pulse regulation in bacterial stress response. Science. 2011;334:366–369. doi: 10.1126/science.1208144.
    1. Starczynowski DT, et al. Identification of miR-145 and miR-146a as mediators of the 5q- syndrome phenotype. Nat. Med. 2010;16:49–58. doi: 10.1038/nm.2054.
    1. Lin SL, Chiang A, Chang D, Ying SY. Loss of mir-146a function in hormone-refractory prostate cancer. RNA. 2008;14:417–424. doi: 10.1261/rna.874808.
    1. Shen J, et al. A functional polymorphism in the miR-146a gene and age of familial breast/ovarian cancer diagnosis. Carcinogenesis. 2008;29:1963–1966. doi: 10.1093/carcin/bgn172.
    1. Ren YG, Zhou XM, Cui ZG, Hou G. Effects of common polymorphisms in miR-146a and miR-196a2 on lung cancer susceptibility: a meta-analysis. J. Thorac. Dis. 2016;8:1297–1305. doi: 10.21037/jtd.2016.05.02.
    1. Marcucci G, et al. Clinical role of microRNAs in cytogenetically normal acute myeloid leukemia: miR-155 upregulation independently identifies high-risk patients. J. Clin. Oncol. 2013;31:2086–2093. doi: 10.1200/JCO.2012.45.6228.
    1. van den Berg A, et al. High expression of B-cell receptor inducible gene BIC in all subtypes of Hodgkin lymphoma. Genes Chromosomes Cancer. 2003;37:20–28. doi: 10.1002/gcc.10186.
    1. Kluiver J, et al. BIC and miR-155 are highly expressed in Hodgkin, primary mediastinal and diffuse large B cell lymphomas. J. Pathol. 2005;207:243–249. doi: 10.1002/path.1825.
    1. Junker A, et al. MicroRNA profiling of multiple sclerosis lesions identifies modulators of the regulatory protein CD47. Brain. 2009;132:3342–3352. doi: 10.1093/brain/awp300.
    1. Hill JM, Pogue AI, Lukiw WJ. Pathogenic microRNAs common to brain and retinal degeneration; recent observations in Alzheimer’s disease and age-related macular degeneration. Front. Neurol. 2015;6:232.
    1. Yang L, et al. miR-146a controls the resolution of T cell responses in mice. J. Exp. Med. 2012;209:1655–1670. doi: 10.1084/jem.20112218.
    1. Lu LF, et al. Function of miR-146a in controlling Treg cell-mediated regulation of Th1 responses. Cell. 2010;142:914–929. doi: 10.1016/j.cell.2010.08.012.
    1. O’Connell RM, Taganov KD, Boldin MP, Cheng G, Baltimore D. MicroRNA-155 is induced during the macrophage inflammatory response. Proc. Natl Acad. Sci. USA. 2007;104:1604–1609. doi: 10.1073/pnas.0610731104.
    1. Doxaki C, Kampranis SC, Eliopoulos AG, Spilianakis C, Tsatsanis C. Coordinated regulation of miR-155 and miR-146a genes during induction of endotoxin tolerance in macrophages. J. Immunol. 2015;195:5750–5761. doi: 10.4049/jimmunol.1500615.
    1. Gantier MP, et al. Analysis of microRNA turnover in mammalian cells following Dicer1 ablation. Nucleic Acids Res. 2011;39:5692–5703. doi: 10.1093/nar/gkr148.
    1. Hao S, Baltimore D. The stability of mRNA influences the temporal order of the induction of genes encoding inflammatory molecules. Nat. Immunol. 2009;10:281–288. doi: 10.1038/ni.1699.
    1. Li Y, et al. miR-146a suppresses invasion of pancreatic cancer cells. Cancer Res. 2010;70:1486–1495. doi: 10.1158/0008-5472.CAN-09-2792.
    1. O’Connell RM, et al. MicroRNAs enriched in hematopoietic stem cells differentially regulate long-term hematopoietic output. Proc. Natl Acad. Sci. USA. 2010;107:14235–14240. doi: 10.1073/pnas.1009798107.
    1. Langmead B, Trapnell C, Pop M, Salzberg SL. Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biol. 2009;10:R25. doi: 10.1186/gb-2009-10-3-r25.
    1. Roberts A, Pachter L. Streaming fragment assignment for real-time analysis of sequencing experiments. Nat. Methods. 2013;10:71–73. doi: 10.1038/nmeth.2251.
    1. Anders S, Huber W. Differential expression analysis for sequence count data. Genome Biol. 2010;11:R106. doi: 10.1186/gb-2010-11-10-r106.

Source: PubMed

3
購読する