Kinase inhibition in autoimmunity and inflammation

Ali A Zarrin, Katherine Bao, Patrick Lupardus, Domagoj Vucic, Ali A Zarrin, Katherine Bao, Patrick Lupardus, Domagoj Vucic

Abstract

Despite recent advances in the treatment of autoimmune and inflammatory diseases, unmet medical needs in some areas still exist. One of the main therapeutic approaches to alleviate dysregulated inflammation has been to target the activity of kinases that regulate production of inflammatory mediators. Small-molecule kinase inhibitors have the potential for broad efficacy, convenience and tissue penetrance, and thus often offer important advantages over biologics. However, designing kinase inhibitors with target selectivity and minimal off-target effects can be challenging. Nevertheless, immense progress has been made in advancing kinase inhibitors with desirable drug-like properties into the clinic, including inhibitors of JAKs, IRAK4, RIPKs, BTK, SYK and TPL2. This Review will address the latest discoveries around kinase inhibitors with an emphasis on clinically validated autoimmunity and inflammatory pathways.

Conflict of interest statement

A.A.Z. is an employee of TRexBio and holds stock in TRexBio and the Roche Group. K.B. is an employee of Genentech. P.L. is an employee of Synthekine and holds stock in Synthekine and the Roche Group. D.V. is an employee of Genentech and holds stock and options in the Roche Group.

Figures

Fig. 1. Current landscape of major druggable…
Fig. 1. Current landscape of major druggable inflammatory receptors and corresponding kinases implicated in human disease.
Major inflammatory pathways and downstream kinases are depicted to show surface versus intracellular drug targets, highlighting drugs that are currently being clinically evaluated or already approved. All biologics that are included have mostly been evaluated in phase II trials and beyond. All small-molecule kinase inhibitors have been evaluated in phase I and beyond. BAFFR, B cell activating-factor receptor; BCMA, B cell maturation antigen; BCR, B cell receptor; BTK, Bruton’s tyrosine kinase; CD40L, CD40 ligand; CSF1R, colony-stimulating factor 1 receptor; CTLA4, cytotoxic T lymphocyte-associated protein 4; FcεR, Fcε receptor; IKKε, inhibitor of NF-κB subunit-ε; IL-1R, IL-1 receptor; IRAK, IL-1R-associated kinase; ITK, IL-2-inducible T cell kinase; JAK, Janus-associated kinase; MD2, myeloid differentiation factor 2; NF-κB, nuclear factor-κ-light-chain-enhancer of activated B cells; NIK, NF-κB-inducing kinase; RANKL, receptor activator of NF-κB ligand; RIP1, receptor-interacting protein 1; RLK, resting lymphocyte kinase; ST2, IL-1R-like 1; SYK, spleen tyrosine kinase; TACI, transmembrane activator and CAML interactor; TBK1, TANK-binding kinase 1; TCR, T cell receptor; TEC, Tec protein tyrosine kinase; TLR, Toll-like receptor; TNF, tumour necrosis factor; TNFR, TNF receptor; TPL2, tumour progression locus 2; TSLP, thymic stromal lymphopoietin; TYK2, tyrosine kinase 2.
Fig. 2. JAK1, JAK2, JAK3 and TYK2…
Fig. 2. JAK1, JAK2, JAK3 and TYK2 integrate the signalling cascades of a diverse set of cytokine and growth receptors.
The binding of extracellular ligands leads to pathway activation via changes to the receptors that permit trans-phosphorylation of associated Janus-associated kinases (JAKs). Activated JAKs then phosphorylate both the receptor and cognate signal transducer and activator of transcription (STAT) proteins. Activated and dimerized STATs then enter the nucleus to bind to transcriptional regulatory sites of target genes. Receptors that use JAK2 and JAK3, JAK3 alone, tyrosine kinase 2 (TYK2) alone or JAK3 and TYK2 have not been described. EPO, erythropoietin; GH, growth hormone; GM-CSF, granulocyte–macrophage colony-stimulating factor; LIF, leukaemia inhibitory factor; OSM, oncostatin M; P, phosphorus; TPO, thrombopoietin; TSLP, thymic stromal lymphopoietin.
Fig. 3. IRAK4 is the upstream kinase…
Fig. 3. IRAK4 is the upstream kinase that transduces TLRs and IL-1R signals.
IL-1 receptor (IL-1R)-associated kinase 1 (IRAK1) and IRAK4 function is regulated by protein–protein interactions and by post-translational modifications that may be uniquely regulated in different cell types to fine-tune an immune response. IL-1R or Toll-like receptor (TLR; except for TLR3) engagement causes the recruitment of adaptor protein myeloid differentiation primary response 88 (MyD88) to the intracellular Toll/IL-1R receptor (TIR) domains to initiate the Myddosome assembly. MyD88 recruits IRAK4 via death domain interactions. IRAK4 is activated via autophosphorylation and is also K63-ubiquitylated (grey circles). Phosphorylated IRAK4 can activate IRAK1, which facilitates IRAK1–tumour necrosis factor receptor-associated factor 6 (TRAF6) complex formation. K48 ubiquitylation (blue circles) of IRAK1 is required for the activation of TGFβ-activated kinase 1-binding protein 1 (TAK1), and its binding to TAB1 and TAB2 to drive the formation of the inhibitor of NF-κB (IKK) complex and subsequent NF-κB inhibitor-α (IκBα) activation, which then leads to NF-κB, mitogen-activated protein kinase (MAPK) and interferon-regulatory factor (IRF) activation to induce the transcription of pro-inflammatory cytokines and cellular processes, such as proliferation and activation,. The mechanism of IRF activation by MyD88 is less understood. Downstream of IRAK4 activation, TAK1 and IKKβ complex can mediate IRF5 phosphorylation, nuclear translocation and transcription in monocytes, but IRAK1 can also directly bind and phosphorylate IRF7 in plasmacytoid dendritic cells. The mechanism of IRAK4 kinase activity-independent action is less understood but may involve K63-ubiquitylated signalling hubs and other novel molecular scaffolds. AP-1, activator protein 1; CREB, cAMP response element-binding protein; P, phosphorus; TAB, TGFβ-activated kinase 1 binding protein; Ub, ubiquitin.
Fig. 4. RIP kinases regulate cell death…
Fig. 4. RIP kinases regulate cell death and inflammatory pathways.
Tumour necrosis factor (TNF) signalling can lead to receptor-interacting protein 1 (RIP1)-dependent apoptosis (mediated by caspase 8) or necroptosis (mediated by RIP3 and mixed-lineage kinase domain-like protein (MLKL)) to cause tissue damage and an inflammatory milieu. RIP1 inhibition can block RIP1-mediated apoptosis and necroptosis, and reduce inflammation by inhibiting inflammatory cell death. The kinase domain of RIP2 allows the binding of E3 ligase X-linked inhibitor of apoptosis protein (XIAP) and subsequent RIP2 ubiquitylation, which is a critical mediator of nucleotide-binding oligomerization domain-containing protein 2 (NOD2) inflammatory signalling. Consequently, RIP2 kinase inhibitors prevent XIAP binding and RIP2 ubiquitylation to inhibit NOD2 pathway-activated NF-κB and mitogen-activated protein kinase (MAPK) signalling, and consequent production and release of pro-inflammatory cytokines, thus blocking inflammation. FADD, FAS-associated death domain; IAP1/2, inhibitor of apoptosis 1 and 2; LUBAC, linear ubiquitin chain assembly complex; MDP, muramyl dipeptide; P, phosphorus; TNFR1, TNF receptor 1; TRADD, TNFR-associated death domain; TRAF2, TNFR-associated factor 2; Ub, ubiquitin.
Fig. 5. ITK and BTK in antigen…
Fig. 5. ITK and BTK in antigen receptor, TLR and FcR signalling.
a | In T cells, IL-2 inducible T cell kinase (ITK) phosphorylates secondary messengers to activate nuclear factor of activated T cells, cytoplasmic 1 (NFATc) and calcium signalling to positively regulate cytokines (such as IL-2) or negatively regulate certain genes such as FASLG. Following T cell receptor (TCR) engagement, active lymphocyte-specific protein tyrosine kinase (LCK) phosphorylates the immunoreceptor tyrosine-based activation motifs (ITAMs), Zap70 and IL-2-inducible T cell kinase (ITK). Zap70 phosphorylates SLP76 and linker for activation of T cells (LAT); ITK phosphorylates phospholipase Cγ1 (PLCγ1). Activated PLCγ1 then hydrolyses phosphatidylinositol 4,5-bisphosphate (PIP2) to produce diacylglcycerol (DAG) and inositol 1,4,5-trisphosphate (IP3), leading to increased calcium flux. b | Bruton’s tyrosine kinase (BTK) regulates multiple receptors including B cell receptor (BCR) or Toll-like receptors (TLRs) in B cells, Fcγ receptor (FCγR) or TLRs in macrophages or plasmacytoid dendritic cells (pDCs) and FCγR or TLRs in basophils and mast cells. Upon BCR engagement, BTK is activated by Src kinases and phosphorylates PLCγ2. PLCγ2 activates NFAT and enhances Ca2+ flux, and activates mitogen-activated protein kinase (MAPK) and NF-κB. BTK has also been implicated in TLR, IC and FCγR or FCεR signalling in multiple cell types via Ca2+ or as yet unknown pathways. AICD, activation-induced cell death; AP-1, activator protein 1; FasL, Fas ligand; IC, immune complex; IRAK4, IL-1 receptor-associated kinase 4; LYN, Lck/Yes-related novel tyrosine kinase; MHC, major histocompatibility complex; MyD88, myeloid differentiation primary response 88; P, phosphorus; PIP3, phosphatidylinositol 3,4,5-trisphosphate; PKC, protein kinase C; SYK, spleen tyrosine kinase.
Fig. 6. TPL2 regulatory inflammatory response downstream…
Fig. 6. TPL2 regulatory inflammatory response downstream of TLRs, TNFR and IL-1R.
The action of both mitogen-activated protein kinase (MAPK) and NF-κB orchestrates the transcription of target genes. In the resting state, p105 prevents and masks tumour progression locus 2 (TPL2) kinase effector function. Agonist stimulation activates the inhibitor of NF-κB (IKK)/NF-κB essential modulator (NEMO) complex. Subsequently, IKKβ phosphorylates IKKα, targeting IKKα for proteasomal degradation. Then, released RelA/p50 dimers are translocated to the nucleus and modulate target gene expression. IKKβ also phosphorylates the target residues S927 and S932 in p105, leading to its proteasomal degradation that results in TPL2 liberation. IKKβ phosphorylates TPL2 at residue S400 to enhance its kinase activity. Free TPL2 then activates MEK1 and MEK2 as well as MEK3 and MEK6 to positively regulate ERK1/2 or p38α/δ to regulate gene transcription via cAMP response element-binding protein (CREB)/activator protein 1 (AP-1) as well as mRNA stability and protein production. The function of A20-binding inhibitor of NF-κB (ABIN2) is not completely understood and involves regulation of protaglandin E2 (PGE2) and cyclooxygenase 2 (COX2) in fibroblasts. TPL2 binds to a small fraction of p105 but the p50 domain of processed p105 can directly impact gene transcription. Post-transcriptional regulation of cytokines and chemokines by MAPKs involves AU-rich elements (AREs) on messenger RNAs to dictate their stability (in the case of tumour necrosis factor (TNF) or IL-6, for instance) or cellular localization (TNF). In addition, other regulatory processes such as CAP-dependent RNA translation and protein export via TNFα-converting enzyme (TACE) (in the case of TNF) are regulated by the TPL2/MAPK axis. Therefore, the net effect of TPL2 inhibition has a profound effect on inflammatory outputs without compromising the NF-κB pathway. IL-1R, IL-1 receptor; P, phosphorus; TLR, Toll-like receptor; TNFR, TNF receptor; Ub, ubiquitin.
Fig. 7. IKKε and TBK1 kinases integrate…
Fig. 7. IKKε and TBK1 kinases integrate signalling from nucleic acid sensors.
Nucleic acid sensors include the endosomal Toll-like receptors (TLRs) TLR3, TLR7, TLR8 and TLR9; the cytosolic DNA sensor cyclic GMP–AMP synthase (cGAS); and retinoic acid-inducible gene I (RIG-I)-like receptors (RLRs). RLRs are RNA sensors that include RIG-I and melanoma differentiation-associated protein 5 (MDA5). Unusual RNAs (double-strand RNA (dsRNA), 5′-phosphorylated (5′P) mRNA) can activate RIG-I or MDA5. Multimerized RIG-I and MDA5 can then bind to create a mitochondrial antiviral-signalling protein (MAVS), mitochondrial-associated membranes and peroxisomes, which in turn activate TANK-binding kinase 1 (TBK1) and inhibitor of NF-κB subunit-ε (IKKε) to activate interferon-regulatory factor 3 (IRF3) and IRF7. Double-stranded DNA (dsDNA; such as self-DNA) can induce an allosteric structural change in cGAS that, in turn, activates second messengers to promote stimulator of interferon genes (STING) to undergo dimerization to form a complex with TBK1 and IKKε that phosphorylates IRF3 to activate gene transcription. DNA-dependent activator of interferon-regulatory factors (DAI; also known as Z-DNA-binding protein 1 (ZBP1)) can also act as a sensor of Z-dsDNA (left-handed double-helical structure), often acquired during viral infection in a cell type-specific manner. DAI recruits TBK1 and IRF3 upon DNA binding and may get further phosphorylated to amplify the DAI/TBK1/IRF3 circuit. Numb-associated kinases (adaptor protein 2 (AP2)-associated protein kinase 1 (AAK1) and cyclin G-associated kinase (GAK)) regulate intracellular viral trafficking during entry, assembly and release of unrelated viruses. Disruption of endocytosis by inhibiting AAK1/GAK can prevent the virus passage into cells. Endosomal TLRs (TLR7/8/9) recruit myeloid differentiation primary response 88 (MyD88) and signal through IL-1 receptor-associated kinase 4 (IRAK4) to activate IRFs similar to cell surface TLRs (see Fig. 3). TIR-domain-containing adapter-inducing IFNβ (TRIF)-dependent and MyD88-independent signalling could also activate TBK1/IKKε to activate IRF-mediated gene transcription. P, phosphorus; ssDNA, single-strand DNA.
Fig. 8. Multiple immune receptors trigger NF-κB…
Fig. 8. Multiple immune receptors trigger NF-κB canonical and non-canonical pathways.
NF-κB-inducing kinase (NIK) itself is regulated at the basal level by a destruction complex, and signal-induced non-canonical NF-κB signalling involves NIK stabilization. The canonical NF-κB pathway involves activation of inhibitor of NF-κB (IKK) complex by TGFβ-activated kinase 1-binding protein 1 (TAK1), IKK-mediated NF-κB inhibitor-α (IκBα) phosphorylation and subsequent degradation, resulting in rapid and transient nuclear translocation of the NF-κB heterodimer RelA/p50. NIK protein levels remain low via active degradation using ubiquitin ligase complex, comprising tumour necrosis factor receptor-associated factor 3 (TRAF3), TRAF2 and inhibitor of apoptosis 1 and 2 (IAP1/2). Receptor activation by agonists recruits this complex to the receptor where activated IAP mediates K48 ubiquitylation and proteasomal degradation of TRAF3, resulting in stabilization and accumulation of NIK. Subsequently, NIK activates IKKα to trigger p100 phosphorylation and processing to enforce persistent activation of RelB/p52 complex to activate gene transcription. BAFF, B cell-activating factor; BCR, B cell receptor; LTβ, lymphotoxin-β; P, phosphorus; TCR, T cell receptor; TLR, Toll-like receptor; TWEAK, tumour necrosis factor-related weak inducer of apoptosis; Ub, ubiquitin.

References

    1. Fullerton JN, Gilroy DW. Resolution of inflammation: a new therapeutic frontier. Nat. Rev. Drug Discov. 2016;15:551–567.
    1. Matthay MA, et al. Acute respiratory distress syndrome. Nat. Rev. Dis. Prim. 2019;5:18.
    1. Spinelli FR, Conti F, Gadina M. HiJAKing SARS-CoV-2? The potential role of JAK inhibitors in the management of COVID-19. Sci. Immunol. 2020;5:eabc5367.
    1. Patterson H, Nibbs R, McInnes I, Siebert S. Protein kinase inhibitors in the treatment of inflammatory and autoimmune diseases. Clin. Exp. Immunol. 2014;176:1–10.
    1. Wang L, Wang FS, Gershwin ME. Human autoimmune diseases: a comprehensive update. J. Intern. Med. 2015;278:369–395.
    1. Krainer J, Siebenhandl S, Weinhausel A. Systemic autoinflammatory diseases. J. Autoimmun. 2020;109:102421.
    1. Paluch C, Santos AM, Anzilotti C, Cornall RJ, Davis SJ. Immune checkpoints as therapeutic targets in autoimmunity. Front. Immunol. 2018;9:2306.
    1. Radawski C, et al. Patient perceptions of unmet medical need in rheumatoid arthritis: a cross-sectional survey in the USA. Rheumatol. Ther. 2019;6:461–471.
    1. Ferguson FM, Gray NS. Kinase inhibitors: the road ahead. Nat. Rev. Drug Discov. 2018;17:353–377.
    1. Cohen P. Targeting protein kinases for the development of anti-inflammatory drugs. Curr. Opin. Cell Biol. 2009;21:317–324.
    1. Lin W, et al. Dual B cell immunotherapy is superior to individual anti-CD20 depletion or BAFF blockade in murine models of spontaneous or accelerated lupus. Arthritis Rheumatol. 2015;67:215–224.
    1. Boleto G, Kanagaratnam L, Drame M, Salmon JH. Safety of combination therapy with two bDMARDs in patients with rheumatoid arthritis: a systematic review and meta-analysis. Semin. Arthritis Rheum. 2018;49:35–42.
    1. Schwartz DM, et al. JAK inhibition as a therapeutic strategy for immune and inflammatory diseases. Nat. Rev. Drug Discov. 2017;16:843–862.
    1. Medzhitov R, Horng T. Transcriptional control of the inflammatory response. Nat. Rev. Immunol. 2009;9:692–703.
    1. Sun SC. The non-canonical NF-κB pathway in immunity and inflammation. Nat. Rev. Immunol. 2017;17:545–558.
    1. Arthur JS, Ley SC. Mitogen-activated protein kinases in innate immunity. Nat. Rev. Immunol. 2013;13:679–692.
    1. Farber DL, Netea MG, Radbruch A, Rajewsky K, Zinkernagel RM. Immunological memory: lessons from the past and a look to the future. Nat. Rev. Immunol. 2016;16:124–128.
    1. Bluestone JA, Mackay CR, O’Shea JJ, Stockinger B. The functional plasticity of T cell subsets. Nat. Rev. Immunol. 2009;9:811–816.
    1. Boothby IC, Cohen JN, Rosenblum MD. Regulatory T cells in skin injury: at the crossroads of tolerance and tissue repair. Sci. Immunol. 2020 doi: 10.1126/sciimmunol.aaz9631.
    1. Adler LN, et al. The other function: class II-restricted antigen presentation by B cells. Front. Immunol. 2017;8:319.
    1. Netea MG, Schlitzer A, Placek K, Joosten LAB, Schultze JL. Innate and adaptive immune memory: an evolutionary continuum in the host’s response to pathogens. Cell Host Microbe. 2019;25:13–26.
    1. Vanamee ES, Faustman DL. Structural principles of tumor necrosis factor superfamily signaling. Sci. Signal. 2018 doi: 10.1126/scisignal.aao4910.
    1. Boraschi D, Italiani P, Weil S, Martin MU. The family of the interleukin-1 receptors. Immunol. Rev. 2018;281:197–232.
    1. Scheller J, Chalaris A, Schmidt-Arras D, Rose-John S. The pro- and anti-inflammatory properties of the cytokine interleukin-6. Biochim. Biophys. Acta. 2011;1813:878–888.
    1. Wojno ED, Hunter CA. New directions in the basic and translational biology of interleukin-27. Trends Immunol. 2012;33:91–97.
    1. Hsieh CS, et al. Development of TH1 CD4+ T cells through IL-12 produced by Listeria-induced macrophages. Science. 1993;260:547–549.
    1. Cua DJ, et al. Interleukin-23 rather than interleukin-12 is the critical cytokine for autoimmune inflammation of the brain. Nature. 2003;421:744–748.
    1. Teng MW, et al. IL-12 and IL-23 cytokines: from discovery to targeted therapies for immune-mediated inflammatory diseases. Nat. Med. 2015;21:719–729.
    1. Clynes R, et al. Modulation of immune complex-induced inflammation in vivo by the coordinate expression of activation and inhibitory Fc receptors. J. Exp. Med. 1999;189:179–185.
    1. Takeuchi O, Akira S. Pattern recognition receptors and inflammation. Cell. 2010;140:805–820.
    1. Senger K, et al. The kinase TPL2 activates ERK and p38 signaling to promote neutrophilic inflammation. Sci Signal. 2017;10:eaah4273.
    1. Kung JE, Jura N. Structural basis for the non-catalytic functions of protein kinases. Structure. 2016;24:7–24.
    1. Rauch J, Volinsky N, Romano D, Kolch W. The secret life of kinases: functions beyond catalysis. Cell Commun. Signal. 2011;9:23.
    1. Danto, S. I. et al. Efficacy and safety of the selective interleukin-1 receptor associated kinase 4 inhibitor, PF-06650833, in patients with active rheumatoid arthritis and inadequate response to methotrexate [Abstract 2909]. Arthritis Rheumatol.70 (Suppl. 10) (2019).
    1. Owen C, Berinstein NL, Christofides A, Sehn LH. Review of Bruton tyrosine kinase inhibitors for the treatment of relapsed or refractory mantle cell lymphoma. Curr. Oncol. 2019;26:e233–e240.
    1. Warr, M. et al. GS-4875, a first-in-class TPL2 inhibitor suppresses MEK-ERK inflammatory signaling and proinflammatory cytokine production in primary human monocytes [Abstract 33]. Arthritis Rheumatol.71 (Suppl. 10) (2019). This is the first report of clinical application of TPL2 inhibitors in inflammatory diseases.
    1. Mullard A. FDA approves first-in-class SYK inhibitor. Nat. Rev. Drug Discov. 2018;17:385.
    1. Burke JR, et al. Autoimmune pathways in mice and humans are blocked by pharmacological stabilization of the TYK2 pseudokinase domain. Sci. Transl Med. 2019 doi: 10.1126/scitranslmed.aaw1736.
    1. Mayence A, Vanden Eynde JJ. Baricitinib: a 2018 novel FDA-approved small molecule inhibiting janus kinases. Pharmaceuticals. 2019 doi: 10.3390/ph12010037.
    1. Xu H, et al. PF-06651600, a dual JAK3/TEC family kinase inhibitor. ACS Chem. Biol. 2019;14:1235–1242.
    1. Tarrant JM, et al. Filgotinib, a JAK1 inhibitor, modulates disease-related biomarkers in rheumatoid arthritis: results from two randomized, controlled phase 2b trials. Rheumatol. Ther. 2020;7:173–190.
    1. Changelian PS, et al. Prevention of organ allograft rejection by a specific Janus kinase 3 inhibitor. Science. 2003;302:875–878.
    1. Kudlacz E, Conklyn M, Andresen C, Whitney-Pickett C, Changelian P. The JAK-3 inhibitor CP-690550 is a potent anti-inflammatory agent in a murine model of pulmonary eosinophilia. Eur. J. Pharmacol. 2008;582:154–161.
    1. Kudlacz E, et al. The novel JAK-3 inhibitor CP-690550 is a potent immunosuppressive agent in various murine models. Am. J. Transpl. 2004;4:51–57.
    1. Vainchenker W, et al. JAK inhibitors for the treatment of myeloproliferative neoplasms and other disorders. F1000Research. 2018;7:82.
    1. Bewersdorf JP, Jaszczur SM, Afifi S, Zhao JC, Zeidan AM. Beyond ruxolitinib: fedratinib and other emergent treatment options for myelofibrosis. Cancer Manag. Res. 2019;11:10777–10790.
    1. Corren J, et al. Tezepelumab in adults with uncontrolled asthma. N. Engl. J. Med. 2017;377:936–946.
    1. Sandborn WJ, et al. Development of gut-selective pan-Janus kinase inhibitor TD-1473 for ulcerative colitis: a translational medicine program. J. Crohns Colitis. 2020 doi: 10.1093/ecco-jcc/jjaa049.
    1. Jones P, et al. Design and synthesis of a pan-Janus kinase inhibitor clinical candidate (PF-06263276) suitable for inhaled and topical delivery for the treatment of inflammatory diseases of the lungs and skin. J. Med. Chem. 2017;60:767–786.
    1. Dengler HS, et al. Lung-restricted inhibition of Janus kinase 1 is effective in rodent models of asthma. Sci. Transl Med. 2018 doi: 10.1126/scitranslmed.aao2151.
    1. Dendrou CA, et al. Resolving TYK2 locus genotype-to-phenotype differences in autoimmunity. Sci. Transl Med. 2016;8:363ra149.
    1. Couturier N, et al. Tyrosine kinase 2 variant influences T lymphocyte polarization and multiple sclerosis susceptibility. Brain. 2011;134:693–703.
    1. Papp K, et al. Phase 2 trial of selective tyrosine kinase 2 inhibition in psoriasis. N. Engl. J. Med. 2018;379:1313–1321.
    1. Richardson P, et al. Baricitinib as potential treatment for 2019-nCoV acute respiratory disease. Lancet. 2020;395:e30–e31.
    1. Cantini F, et al. Baricitinib therapy in COVID-19: a pilot study on safety and clinical impact. J. Infect. 2020 doi: 10.1016/j.jinf.2020.04.017.
    1. Motshwene PG, et al. An oligomeric signaling platform formed by the Toll-like receptor signal transducers MyD88 and IRAK-4. J. Biol. Chem. 2009;284:25404–25411.
    1. Lin SC, Lo YC, Wu H. Helical assembly in the MyD88–IRAK4–IRAK2 complex in TLR/IL-1R signalling. Nature. 2010;465:885–890.
    1. Ferrao R, et al. IRAK4 dimerization and trans-autophosphorylation are induced by Myddosome assembly. Mol. Cell. 2014;55:891–903.
    1. Kawagoe T, et al. Sequential control of Toll-like receptor-dependent responses by IRAK1 and IRAK2. Nat. Immunol. 2008;9:684–691.
    1. Li S, Strelow A, Fontana EJ, Wesche H. IRAK-4: a novel member of the IRAK family with the properties of an IRAK-kinase. Proc. Natl Acad. Sci. USA. 2002;99:5567–5572.
    1. Cao Z, Xiong J, Takeuchi M, Kurama T, Goeddel DV. TRAF6 is a signal transducer for interleukin-1. Nature. 1996;383:443–446.
    1. Jiang Z, Ninomiya-Tsuji J, Qian Y, Matsumoto K, Li X. Interleukin-1 (IL-1) receptor-associated kinase-dependent IL-1-induced signaling complexes phosphorylate TAK1 and TAB2 at the plasma membrane and activate TAK1 in the cytosol. Mol. Cell Biol. 2002;22:7158–7167.
    1. Wan Y, et al. Interleukin-1 receptor-associated kinase 2 is critical for lipopolysaccharide-mediated post-transcriptional control. J. Biol. Chem. 2009;284:10367–10375.
    1. Ye H, et al. Distinct molecular mechanism for initiating TRAF6 signalling. Nature. 2002;418:443–447.
    1. Zhou H, et al. IRAK2 directs stimulus-dependent nuclear export of inflammatory mRNAs. eLife. 2017 doi: 10.7554/eLife.29630.
    1. Wesche H, et al. IRAK-M is a novel member of the Pelle/interleukin-1 receptor-associated kinase (IRAK) family. J. Biol. Chem. 1999;274:19403–19410.
    1. Kobayashi K, et al. IRAK-M is a negative regulator of Toll-like receptor signaling. Cell. 2002;110:191–202.
    1. Zhou H, et al. IRAK-M mediates Toll-like receptor/IL-1R-induced NFκB activation and cytokine production. EMBO J. 2013;32:583–596.
    1. Su J, Zhang T, Tyson J, Li L. The interleukin-1 receptor-associated kinase M selectively inhibits the alternative, instead of the classical NFκB pathway. J. Innate Immun. 2009;1:164–174.
    1. Chiang CY, et al. Cofactors required for TLR7- and TLR9-dependent innate immune responses. Cell Host Microbe. 2012;11:306–318.
    1. Negishi H, et al. Evidence for licensing of IFN-γ-induced IFN regulatory factor 1 transcription factor by MyD88 in Toll-like receptor-dependent gene induction program. Proc. Natl Acad. Sci. USA. 2006;103:15136–15141.
    1. Schmitz F, et al. Interferon-regulatory-factor 1 controls Toll-like receptor 9-mediated IFN-β production in myeloid dendritic cells. Eur. J. Immunol. 2007;37:315–327.
    1. Ku CL, et al. IRAK4 and NEMO mutations in otherwise healthy children with recurrent invasive pneumococcal disease. J. Med. Genet. 2007;44:16–23.
    1. Cushing L, et al. IRAK4 kinase activity controls Toll-like receptor-induced inflammation through the transcription factor IRF5 in primary human monocytes. J. Biol. Chem. 2017;292:18689–18698.
    1. Kawagoe T, et al. Essential role of IRAK-4 protein and its kinase activity in Toll-like receptor-mediated immune responses but not in TCR signaling. J. Exp. Med. 2007;204:1013–1024.
    1. Swantek JL, Tsen MF, Cobb MH, Thomas JA. IL-1 receptor-associated kinase modulates host responsiveness to endotoxin. J. Immunol. 2000;164:4301–4306.
    1. Koziczak-Holbro M, et al. IRAK-4 kinase activity is required for interleukin-1 (IL-1) receptor- and Toll-like receptor 7-mediated signaling and gene expression. J. Biol. Chem. 2007;282:13552–13560.
    1. Dudhgaonkar S, et al. Selective IRAK4 inhibition attenuates disease in murine lupus models and demonstrates steroid sparing activity. J. Immunol. 2017;198:1308–1319.
    1. Murphy M, Pattabiraman G, Manavalan TT, Medvedev AE. Deficiency in IRAK4 activity attenuates manifestations of murine lupus. Eur. J. Immunol. 2017;47:880–891.
    1. Nanda SK, Lopez-Pelaez M, Arthur JS, Marchesi F, Cohen P. Suppression of IRAK1 or IRAK4 catalytic activity, but not type 1 IFN signaling, prevents lupus nephritis in mice expressing a ubiquitin binding-defective mutant of ABIN1. J. Immunol. 2016;197:4266–4273.
    1. Kim SH, et al. The dietary flavonoid kaempferol mediates anti-inflammatory responses via the Src, Syk, IRAK1, and IRAK4 molecular targets. Mediators Inflamm. 2015;2015:904142.
    1. Rekhter M, et al. Genetic ablation of IRAK4 kinase activity inhibits vascular lesion formation. Biochem. Biophys. Res. Commun. 2008;367:642–648.
    1. Cameron B, et al. Loss of interleukin receptor-associated kinase 4 signaling suppresses amyloid pathology and alters microglial phenotype in a mouse model of Alzheimer’s disease. J. Neurosci. 2012;32:15112–15123.
    1. Thomas JA, et al. Impaired cytokine signaling in mice lacking the IL-1 receptor-associated kinase. J. Immunol. 1999;163:978–984.
    1. Della Mina E, et al. Inherited human IRAK-1 deficiency selectively impairs TLR signaling in fibroblasts. Proc. Natl Acad. Sci. USA. 2017;114:E514–E523.
    1. Sun J, et al. Comprehensive RNAi-based screening of human and mouse TLR pathways identifies species-specific preferences in signaling protein use. Sci. Signal. 2016;9:ra3.
    1. Perkins DJ, Vogel SN. Inflammation: species-specific TLR signalling — insight into human disease. Nat. Rev. Rheumatol. 2016;12:198–200.
    1. von Bernuth H, Picard C, Puel A, Casanova JL. Experimental and natural infections in MyD88- and IRAK-4-deficient mice and humans. Eur. J. Immunol. 2012;42:3126–3135.
    1. Pauls E, et al. Two phases of inflammatory mediator production defined by the study of IRAK2 and IRAK1 knock-in mice. J. Immunol. 2013;191:2717–2730.
    1. Flannery SM, Keating SE, Szymak J, Bowie AG. Human interleukin-1 receptor-associated kinase-2 is essential for Toll-like receptor-mediated transcriptional and post-transcriptional regulation of tumor necrosis factor α. J. Biol. Chem. 2011;286:23688–23697.
    1. Alsina L, et al. A narrow repertoire of transcriptional modules responsive to pyogenic bacteria is impaired in patients carrying loss-of-function mutations in MYD88 or IRAK4. Nat. Immunol. 2014;15:1134–1142.
    1. Emmerich CH, et al. Activation of the canonical IKK complex by K63/M1-linked hybrid ubiquitin chains. Proc. Natl Acad. Sci. USA. 2013;110:15247–15252.
    1. De S, et al. Mechanism of dysfunction of human variants of the IRAK4 kinase and a role for its kinase activity in interleukin-1 receptor signaling. J. Biol. Chem. 2018;293:15208–15220.
    1. De Nardo D, et al. Interleukin-1 receptor-associated kinase 4 (IRAK4) plays a dual role in Myddosome formation and Toll-like receptor signaling. J. Biol. Chem. 2018;293:15195–15207.
    1. Silke J, Rickard JA, Gerlic M. The diverse role of RIP kinases in necroptosis and inflammation. Nat. Immunol. 2015;16:689–697.
    1. Witt A, Vucic D. Diverse ubiquitin linkages regulate RIP kinases-mediated inflammatory and cell death signaling. Cell Death Differ. 2017;24:1160–1171.
    1. Newton K. RIPK1 and RIPK3: critical regulators of inflammation and cell death. Trends Cell Biol. 2015;25:347–353.
    1. Varfolomeev E, Vucic D. Intracellular regulation of TNF activity in health and disease. Cytokine. 2018;101:26–32.
    1. Bertrand MJ, et al. cIAP1 and cIAP2 facilitate cancer cell survival by functioning as E3 ligases that promote RIP1 ubiquitination. Mol. Cell. 2008;30:689–700.
    1. Varfolomeev E, et al. c-IAP1 and c-IAP2 are critical mediators of tumor necrosis factor α (TNFα)-induced NF-κB activation. J. Biol. Chem. 2008;283:24295–24299.
    1. Dynek JN, et al. c-IAP1 and UbcH5 promote K11-linked polyubiquitination of RIP1 in TNF signalling. EMBO J. 2010;29:4198–4209.
    1. Newton K. Multitasking kinase RIPK1 regulates cell death and inflammation. Cold Spring Harb. Perspect. Biol. 2019 doi: 10.1101/cshperspect.a036368.
    1. de Almagro MC, et al. Coordinated ubiquitination and phosphorylation of RIP1 regulates necroptotic cell death. Cell Death Differ. 2017;24:26–37.
    1. Lawlor KE, et al. RIPK3 promotes cell death and NLRP3 inflammasome activation in the absence of MLKL. Nat. Commun. 2015;6:6282.
    1. Patel S, et al. RIP1 inhibition blocks inflammatory diseases but not tumor growth or metastases. Cell Death Differ. 2020;27:161–175.
    1. Newton K, et al. Activity of protein kinase RIPK3 determines whether cells die by necroptosis or apoptosis. Science. 2014;343:1357–1360.
    1. Mandal P, et al. RIP3 induces apoptosis independent of pronecrotic kinase activity. Mol. Cell. 2014;56:481–495.
    1. Newton K, et al. RIPK3 deficiency or catalytically inactive RIPK1 provides greater benefit than MLKL deficiency in mouse models of inflammation and tissue injury. Cell Death Differ. 2016;23:1565–1576.
    1. Mulay SR, et al. Cytotoxicity of crystals involves RIPK3–MLKL-mediated necroptosis. Nat. Commun. 2016;7:10274.
    1. Webster, J. D. et al. RIP1 kinase activity is critical for skin inflammation but not for viral propagation. J. Leukoc. Biol. (2020).
    1. Duan X, et al. Inhibition of keratinocyte necroptosis mediated by RIPK1/RIPK3/MLKL provides a protective effect against psoriatic inflammation. Cell Death Dis. 2020;11:134.
    1. Alevy YG, et al. IL-13-induced airway mucus production is attenuated by MAPK13 inhibition. J. Clin. Invest. 2012;122:4555–4568.
    1. Dondelinger Y, et al. NF-κB-independent role of IKKα/IKKβ in preventing RIPK1 kinase-dependent apoptotic and necroptotic cell death during TNF signaling. Mol. Cell. 2015;60:63–76.
    1. Geng J, et al. Regulation of RIPK1 activation by TAK1-mediated phosphorylation dictates apoptosis and necroptosis. Nat. Commun. 2017;8:359.
    1. Tao P, et al. A dominant autoinflammatory disease caused by non-cleavable variants of RIPK1. Nature. 2020;577:109–114.
    1. Newton K, et al. Cleavage of RIPK1 by caspase-8 is crucial for limiting apoptosis and necroptosis. Nature. 2019;574:428–431.
    1. Lalaoui N, et al. Mutations that prevent caspase cleavage of RIPK1 cause autoinflammatory disease. Nature. 2020;577:103–108.
    1. Matsuzawa-Ishimoto Y, et al. Autophagy protein ATG16L1 prevents necroptosis in the intestinal epithelium. J. Exp. Med. 2017;214:3687–3705.
    1. Linkermann A, et al. Rip1 (receptor-interacting protein kinase 1) mediates necroptosis and contributes to renal ischemia/reperfusion injury. Kidney Int. 2012;81:751–761.
    1. Degterev A, Ofengeim D, Yuan J. Targeting RIPK1 for the treatment of human diseases. Proc. Natl Acad. Sci. USA. 2019;116:9714–9722.
    1. Harris PA, et al. Discovery of a first-in-class receptor interacting protein 1 (RIP1) kinase specific clinical candidate (GSK2982772) for the treatment of inflammatory diseases. J. Med. Chem. 2017;60:1247–1261.
    1. Caruso R, Warner N, Inohara N, Nunez G. NOD1 and NOD2: signaling, host defense, and inflammatory disease. Immunity. 2014;41:898–908.
    1. Caso F, et al. Autoinflammatory granulomatous diseases: from Blau syndrome and early-onset sarcoidosis to NOD2-mediated disease and Crohn’s disease. RMD Open. 2015;1:e000097.
    1. Damgaard RB, et al. The ubiquitin ligase XIAP recruits LUBAC for NOD2 signaling in inflammation and innate immunity. Mol. Cell. 2012;46:746–758.
    1. Goncharov T, et al. Disruption of XIAP–RIP2 association blocks NOD2-mediated inflammatory signaling. Mol. Cell. 2018;69:551–565.
    1. Hrdinka M, et al. Small molecule inhibitors reveal an indispensable scaffolding role of RIPK2 in NOD2 signaling. EMBO J. 2018 doi: 10.15252/embj.201899372.
    1. Schwartzberg PL, Finkelstein LD, Readinger JA. TEC-family kinases: regulators of T-helper-cell differentiation. Nat. Rev. Immunol. 2005;5:284–295.
    1. Vargas L, Hamasy A, Nore BF, Smith CI. Inhibitors of BTK and ITK: state of the new drugs for cancer, autoimmunity and inflammatory diseases. Scand. J. Immunol. 2013;78:130–139.
    1. Schaeffer EM, et al. Tec family kinases modulate thresholds for thymocyte development and selection. J. Exp. Med. 2000;192:987–1000.
    1. Sun Y, et al. Inhibition of the kinase ITK in a mouse model of asthma reduces cell death and fails to inhibit the inflammatory response. Sci. Signal. 2015;8:ra122.
    1. Nagata S. Apoptosis by death factor. Cell. 1997;88:355–365.
    1. Long M, et al. Ibrutinib treatment improves T cell number and function in CLL patients. J. Clin. Invest. 2017;127:3052–3064.
    1. Khan WN, et al. Defective B cell development and function in Btk-deficient mice. Immunity. 1995;3:283–299.
    1. Luk ADW, et al. Type I and III interferon productions are impaired in X-linked agammaglobulinemia patients toward poliovirus but not influenza virus. Front. Immunol. 2018;9:1826.
    1. Nimmerjahn F, Ravetch JV. Fcγ receptors as regulators of immune responses. Nat. Rev. Immunol. 2008;8:34–47.
    1. Kawakami Y, et al. Tyrosine phosphorylation and activation of Bruton tyrosine kinase upon Fcε RI cross-linking. Mol. Cell Biol. 1994;14:5108–5113.
    1. Hata D, et al. Involvement of Bruton’s tyrosine kinase in FcεRI-dependent mast cell degranulation and cytokine production. J. Exp. Med. 1998;187:1235–1247.
    1. Smiljkovic D, et al. BTK inhibition is a potent approach to block IgE-mediated histamine release in human basophils. Allergy. 2017;72:1666–1676.
    1. Di Paolo JA, et al. Specific Btk inhibition suppresses B cell- and myeloid cell-mediated arthritis. Nat. Chem. Biol. 2011;7:41–50.
    1. Katewa A, et al. Btk-specific inhibition blocks pathogenic plasma cell signatures and myeloid cell-associated damage in IFNα-driven lupus nephritis. JCI Insight. 2017;2:e90111.
    1. Takata M, Homma Y, Kurosaki T. Requirement of phospholipase C-γ 2 activation in surface immunoglobulin M-induced B cell apoptosis. J. Exp. Med. 1995;182:907–914.
    1. Fluckiger AC, et al. Btk/Tec kinases regulate sustained increases in intracellular Ca2+ following B-cell receptor activation. EMBO J. 1998;17:1973–1985.
    1. Wang D, et al. Phospholipase Cγ2 is essential in the functions of B cell and several Fc receptors. Immunity. 2000;13:25–35.
    1. Wang J, Lau KY, Jung J, Ravindran P, Barrat FJ. Bruton’s tyrosine kinase regulates TLR9 but not TLR7 signaling in human plasmacytoid dendritic cells. Eur. J. Immunol. 2014;44:1130–1136.
    1. Bender AT, et al. Ability of Bruton’s tyrosine kinase inhibitors to sequester Y551 and prevent phosphorylation determines potency for inhibition of fc receptor but not B-cell receptor signaling. Mol. Pharmacol. 2017;91:208–219.
    1. Gillooly KM, et al. Bruton’s tyrosine kinase inhibitor BMS-986142 in experimental models of rheumatoid arthritis enhances efficacy of agents representing clinical standard-of-care. PLoS ONE. 2017;12:e0181782.
    1. Chalmers SA, et al. Therapeutic blockade of immune complex-mediated glomerulonephritis by highly selective inhibition of Bruton’s tyrosine yinase. Sci. Rep. 2016;6:26164.
    1. Bender AT, et al. Btk inhibition treats TLR7/IFN driven murine lupus. Clin. Immunol. 2016;164:65–77.
    1. Corzo CA, et al. The kinase IRAK4 promotes endosomal TLR and immune complex signaling in B cells and plasmacytoid dendritic cells. Sci. Signal. 2020 doi: 10.1126/scisignal.aaz1053.
    1. Montalban X, et al. Placebo-controlled trial of an oral BTK inhibitor in multiple sclerosis. N. Engl. J. Med. 2019;380:2406–2417.
    1. Watterson SH, et al. Discovery of branebrutinib (BMS-986195): a strategy for identifying a highly potent and selective covalent inhibitor providing rapid in vivo unactivation of Bruton’s tyrosine kinase (BTK) J. Med. Chem. 2019;62:3228–3250.
    1. Cohen, S. et al. Efficacy and safety of fenebrutinib, a BTK inhibitor, compared to placebo in rheumatoid arthritis patients with active disease despite TNF inhibitor treatment: randomized, double blind, phase 2 study [Abstract 929]. Arthritis Rheumatol.71 (Suppl. 10) (2019).
    1. Treon SP, et al. The BTK-inhibitor ibrutinib may protect against pulmonary injury in COVID-19 infected patients. Blood. 2020 doi: 10.1182/blood.2020006288.
    1. Mocsai A, Ruland J, Tybulewicz VL. The SYK tyrosine kinase: a crucial player in diverse biological functions. Nat. Rev. Immunol. 2010;10:387–402.
    1. Gradler U, et al. Structural and biophysical characterization of the Syk activation switch. J. Mol. Biol. 2013;425:309–333.
    1. Deindl S, et al. Structural basis for the inhibition of tyrosine kinase activity of ZAP-70. Cell. 2007;129:735–746.
    1. Keshvara LM, Isaacson C, Harrison ML, Geahlen RL. Syk activation and dissociation from the B-cell antigen receptor is mediated by phosphorylation of tyrosine 130. J. Biol. Chem. 1997;272:10377–10381.
    1. Lupher ML, Jr, et al. Cbl-mediated negative regulation of the Syk tyrosine kinase. A critical role for Cbl phosphotyrosine-binding domain binding to Syk phosphotyrosine 323. J. Biol. Chem. 1998;273:35273–35281.
    1. Yankee TM, Keshvara LM, Sawasdikosol S, Harrison ML, Geahlen RL. Inhibition of signaling through the B cell antigen receptor by the protooncogene product, c-Cbl, requires Syk tyrosine 317 and the c-Cbl phosphotyrosine-binding domain. J. Immunol. 1999;163:5827–5835.
    1. Geahlen RL. Getting Syk: spleen tyrosine kinase as a therapeutic target. Trends Pharmacol. Sci. 2014;35:414–422.
    1. Ozaki N, et al. Syk-dependent signaling pathways in neutrophils and macrophages are indispensable in the pathogenesis of anti-collagen antibody-induced arthritis. Int. Immunol. 2012;24:539–550.
    1. Cheng AM, et al. Syk tyrosine kinase required for mouse viability and B-cell development. Nature. 1995;378:303–306.
    1. Turner M, et al. Perinatal lethality and blocked B-cell development in mice lacking the tyrosine kinase Syk. Nature. 1995;378:298–302.
    1. Jakus Z, Simon E, Balazs B, Mocsai A. Genetic deficiency of Syk protects mice from autoantibody-induced arthritis. Arthritis Rheum. 2010;62:1899–1910.
    1. Pine PR, et al. Inflammation and bone erosion are suppressed in models of rheumatoid arthritis following treatment with a novel Syk inhibitor. Clin. Immunol. 2007;124:244–257.
    1. Coffey G, et al. Specific inhibition of spleen tyrosine kinase suppresses leukocyte immune function and inflammation in animal models of rheumatoid arthritis. J. Pharmacol. Exp. Ther. 2012;340:350–359.
    1. Genovese MC, et al. A phase III, multicenter, randomized, double-blind, placebo-controlled, parallel-group study of 2 dosing regimens of fostamatinib in patients with rheumatoid arthritis with an inadequate response to a tumor necrosis factor-alpha antagonist. J. Rheumatol. 2014;41:2120–2128.
    1. Weinblatt ME, et al. Treatment of rheumatoid arthritis with a Syk kinase inhibitor: a twelve-week, randomized, placebo-controlled trial. Arthritis Rheum. 2008;58:3309–3318.
    1. Braselmann S, et al. R406, an orally available spleen tyrosine kinase inhibitor blocks fc receptor signaling and reduces immune complex-mediated inflammation. J. Pharmacol. Exp. Ther. 2006;319:998–1008.
    1. Kunwar S, Devkota AR, Ghimire DK. Fostamatinib, an oral spleen tyrosine kinase inhibitor, in the treatment of rheumatoid arthritis: a meta-analysis of randomized controlled trials. Rheumatol. Int. 2016;36:1077–1087.
    1. Weinblatt ME, et al. An oral spleen tyrosine kinase (Syk) inhibitor for rheumatoid arthritis. N. Engl. J. Med. 2010;363:1303–1312.
    1. Markham A. Fostamatinib: first global approval. Drugs. 2018;78:959–963.
    1. Newland A, Lee EJ, McDonald V, Bussel JB. Fostamatinib for persistent/chronic adult immune thrombocytopenia. Immunotherapy. 2018;10:9–25.
    1. Rolf MG, et al. In vitro pharmacological profiling of R406 identifies molecular targets underlying the clinical effects of fostamatinib. Pharmacol. Res. Perspect. 2015;3:e00175.
    1. Simon M, Vanes L, Geahlen RL, Tybulewicz VL. Distinct roles for the linker region tyrosines of Syk in FcεRI signaling in primary mast cells. J. Biol. Chem. 2005;280:4510–4517.
    1. Chen L, et al. ZAP-70 directly enhances IgM signaling in chronic lymphocytic leukemia. Blood. 2005;105:2036–2041.
    1. Currie KS, et al. Discovery of GS-9973, a selective and orally efficacious inhibitor of spleen tyrosine kinase. J. Med. Chem. 2014;57:3856–3873.
    1. Sharman J, et al. An open-label phase 2 trial of entospletinib (GS-9973), a selective spleen tyrosine kinase inhibitor, in chronic lymphocytic leukemia. Blood. 2015;125:2336–2343.
    1. Barr PM, et al. Phase 2 study of idelalisib and entospletinib: pneumonitis limits combination therapy in relapsed refractory CLL and NHL. Blood. 2016;127:2411–2415.
    1. Xu D, Matsumoto ML, McKenzie BS, Zarrin AA. TPL2 kinase action and control of inflammation. Pharmacol. Res. 2017 doi: 10.1016/j.phrs.2017.11.031.
    1. Beinke S, Robinson MJ, Hugunin M, Ley SC. Lipopolysaccharide activation of the TPL-2/MEK/extracellular signal-regulated kinase mitogen-activated protein kinase cascade is regulated by IκB kinase-induced proteolysis of NF-κB1 p105. Mol. Cell Biol. 2004;24:9658–9667.
    1. Lang V, et al. ABIN-2 forms a ternary complex with TPL-2 and NF-κB1 p105 and is essential for TPL-2 protein stability. Mol. Cell Biol. 2004;24:5235–5248.
    1. Beinke S, et al. NF-κB1 p105 negatively regulates TPL-2 MEK kinase activity. Mol. Cell Biol. 2003;23:4739–4752.
    1. Dumitru CD, et al. TNF-α induction by LPS is regulated posttranscriptionally via a Tpl2/ERK-dependent pathway. Cell. 2000;103:1071–1083.
    1. Das S, et al. Tpl2/cot signals activate ERK, JNK, and NF-κB in a cell-type and stimulus-specific manner. J. Biol. Chem. 2005;280:23748–23757.
    1. Rousseau S, et al. TPL2-mediated activation of ERK1 and ERK2 regulates the processing of pre-TNFα in LPS-stimulated macrophages. J. Cell Sci. 2008;121:149–154.
    1. Waterfield MR, Zhang M, Norman LP, Sun SC. NF-κB1/p105 regulates lipopolysaccharide-stimulated MAP kinase signaling by governing the stability and function of the Tpl2 kinase. Mol. Cell. 2003;11:685–694.
    1. Sriskantharajah S, et al. Regulation of experimental autoimmune encephalomyelitis by TPL-2 kinase. J. Immunol. 2014;192:3518–3529.
    1. Nanda SK, et al. ABIN2 function is required to suppress DSS-induced colitis by a Tpl2-independent mechanism. J. Immunol. 2018;201:3373–3382.
    1. Roulis M, et al. Intestinal myofibroblast-specific Tpl2–Cox-2–PGE2 pathway links innate sensing to epithelial homeostasis. Proc. Natl Acad. Sci. USA. 2014;111:E4658–E4667.
    1. Cuenda A, Rousseau S. p38 MAP-kinases pathway regulation, function and role in human diseases. Biochim. Biophys. Acta. 2007;1773:1358–1375.
    1. Remy G, et al. Differential activation of p38MAPK isoforms by MKK6 and MKK3. Cell Signal. 2010;22:660–667.
    1. Jiang Y, et al. Characterization of the structure and function of the fourth member of p38 group mitogen-activated protein kinases, p38δ. J. Biol. Chem. 1997;272:30122–30128.
    1. Risco A, et al. p38γ and p38δ kinases regulate the Toll-like receptor 4 (TLR4)-induced cytokine production by controlling ERK1/2 protein kinase pathway activation. Proc. Natl Acad. Sci. USA. 2012;109:11200–11205.
    1. Gonzalez-Teran B, et al. Eukaryotic elongation factor 2 controls TNF-α translation in LPS-induced hepatitis. J. Clin. Invest. 2013;123:164–178.
    1. Menon MB, Gaestel M. TPL2 meets p38MAPK: emergence of a novel positive feedback loop in inflammation. Biochem. J. 2016;473:2995–2999.
    1. Criado G, et al. Alternative p38 MAPKs are essential for collagen-induced arthritis. Arthritis Rheumatol. 2014;66:1208–1217.
    1. Ittner A, et al. Regulation of PTEN activity by p38δ–PKD1 signaling in neutrophils confers inflammatory responses in the lung. J. Exp. Med. 2012;209:2229–2246.
    1. Sumara G, et al. Regulation of PKD by the MAPK p38δ in insulin secretion and glucose homeostasis. Cell. 2009;136:235–248.
    1. Tang J, Qi X, Mercola D, Han J, Chen G. Essential role of p38γ in K-Ras transformation independent of phosphorylation. J. Biol. Chem. 2005;280:23910–23917.
    1. Eyers PA, Craxton M, Morrice N, Cohen P, Goedert M. Conversion of SB 203580-insensitive MAP kinase family members to drug-sensitive forms by a single amino-acid substitution. Chem. Biol. 1998;5:321–328.
    1. Fitzgerald CE, et al. Structural basis for p38α MAP kinase quinazolinone and pyridol-pyrimidine inhibitor specificity. Nat. Struct. Biol. 2003;10:764–769.
    1. Gum RJ, et al. Acquisition of sensitivity of stress-activated protein kinases to the p38 inhibitor, SB 203580, by alteration of one or more amino acids within the ATP binding pocket. J. Biol. Chem. 1998;273:15605–15610.
    1. Gangwal RP, Bhadauriya A, Damre MV, Dhoke GV, Sangamwar AT. p38 mitogen-activated protein kinase inhibitors: a review on pharmacophore mapping and QSAR studies. Curr. Top. Medicinal Chem. 2013;13:1015–1035.
    1. Coulthard LR, White DE, Jones DL, McDermott MF, Burchill SA. p38MAPK: stress responses from molecular mechanisms to therapeutics. Trends Mol. Med. 2009;15:369–379.
    1. Wang C, et al. Selective inhibition of the p38α MAPK–MK2 axis inhibits inflammatory cues including inflammasome priming signals. J. Exp. Med. 2018;215:1315–1325.
    1. Schreiber S, et al. Oral p38 mitogen-activated protein kinase inhibition with BIRB 796 for active Crohn’s disease: a randomized, double-blind, placebo-controlled trial. Clin. Gastroenterol. Hepatol. 2006;4:325–334.
    1. Charron CE, et al. RV568, a narrow-spectrum kinase inhibitor with p38 MAPK-α and -γ selectivity, suppresses COPD inflammation. Eur. Respir. J. 2017 doi: 10.1183/13993003.00188-2017.
    1. Armendariz-Borunda J, et al. A controlled clinical trial with pirfenidone in the treatment of pathological skin scarring caused by burns in pediatric patients. Ann. Plastic Surg. 2012;68:22–28.
    1. Margaritopoulos GA, Vasarmidi E, Antoniou KM. Pirfenidone in the treatment of idiopathic pulmonary fibrosis: an evidence-based review of its place in therapy. Core Evid. 2016;11:11–22.
    1. Khanna D, et al. An open-label, phase II study of the safety and tolerability of pirfenidone in patients with scleroderma-associated interstitial lung disease: the LOTUSS trial. J. Rheumatol. 2016;43:1672–1679.
    1. Rodriguez-Castellanos M, Tlacuilo-Parra A, Sanchez-Enriquez S, Velez-Gomez E, Guevara-Gutierrez E. Pirfenidone gel in patients with localized scleroderma: a phase II study. Arthritis Res. Ther. 2015;16:510.
    1. Sharma K, et al. Pirfenidone for diabetic nephropathy. J. Am. Soc. Nephrol. 2011;22:1144–1151.
    1. Flores-Contreras L, et al. Treatment with pirfenidone for two years decreases fibrosis, cytokine levels and enhances CB2 gene expression in patients with chronic hepatitis C. BMC Gastroenterol. 2014;14:131.
    1. Kondoh Y, et al. Comparative chemical array screening for p38γ/δ MAPK inhibitors using a single gatekeeper residue difference between p38α/β and p38γ/δ. Sci. Rep. 2016;6:29881.
    1. Nithianandarajah-Jones GN, Wilm B, Goldring CE, Muller J, Cross MJ. ERK5: structure, regulation and function. Cell Signal. 2012;24:2187–2196.
    1. Lin W, et al. Function of CSF1 and IL34 in macrophage homeostasis, inflammation, and cancer. Front. Immunol. 2019;10:2019.
    1. Kumari A, Silakari O, Singh RK. Recent advances in colony stimulating factor-1 receptor/c-FMS as an emerging target for various therapeutic implications. Biomed. Pharmacother. 2018;103:662–679.
    1. Yan C, Luo H, Lee JD, Abe J, Berk BC. Molecular cloning of mouse ERK5/BMK1 splice variants and characterization of ERK5 functional domains. J. Biol. Chem. 2001;276:10870–10878.
    1. Buschbeck M, Ullrich A. The unique C-terminal tail of the mitogen-activated protein kinase ERK5 regulates its activation and nuclear shuttling. J. Biol. Chem. 2005;280:2659–2667.
    1. Kasler HG, Victoria J, Duramad O, Winoto A. ERK5 is a novel type of mitogen-activated protein kinase containing a transcriptional activation domain. Mol. Cell Biol. 2000;20:8382–8389.
    1. Wilhelmsen K, et al. Extracellular signal-regulated kinase 5 promotes acute cellular and systemic inflammation. Sci. Signal. 2015;8:ra86.
    1. Lin EC, et al. ERK5 kinase activity is dispensable for cellular immune response and proliferation. Proc. Natl Acad. Sci. USA. 2016;113:11865–11870.
    1. Finegan KG, et al. ERK5 is a critical mediator of inflammation-driven cancer. Cancer Res. 2015;75:742–753.
    1. Hayashi M, et al. Targeted deletion of BMK1/ERK5 in adult mice perturbs vascular integrity and leads to endothelial failure. J. Clin. Invest. 2004;113:1138–1148.
    1. Platanias LC. Mechanisms of type-I- and type-II-interferon-mediated signalling. Nat. Rev. Immunol. 2005;5:375–386.
    1. Eleftheriou D, Brogan PA. Genetic interferonopathies: an overview. Best Pract. Res. Clin. Rheumatol. 2017;31:441–459.
    1. Baechler EC, et al. Interferon-inducible gene expression signature in peripheral blood cells of patients with severe lupus. Proc. Natl Acad. Sci. USA. 2003;100:2610–2615.
    1. Volpi S, Picco P, Caorsi R, Candotti F, Gattorno M. Type I interferonopathies in pediatric rheumatology. Pediatric Rheumatol. 2016;14:35.
    1. Reilly SM, et al. An inhibitor of the protein kinases TBK1 and IKK-ε improves obesity-related metabolic dysfunctions in mice. Nat. Med. 2013;19:313–321.
    1. Neveu G, et al. AP-2-associated protein kinase 1 and cyclin G-associated kinase regulate hepatitis C virus entry and are potential drug targets. J. Virol. 2015;89:4387–4404.
    1. Schor S, Einav S. Repurposing of kinase inhibitors as broad-spectrum antiviral drugs. DNA Cell Biol. 2018;37:63–69.
    1. Ahmad L, Zhang SY, Casanova JL, Sancho-Shimizu V. Human TBK1: a gatekeeper of neuroinflammation. Trends Mol. Med. 2016;22:511–527.
    1. Peters RT, Maniatis T. A new family of IKK-related kinases may function as IκB kinase kinases. Biochim. Biophys. Acta. 2001;1471:M57–M62.
    1. Zhang L, Zhao X, Zhang M, Zhao W, Gao C. Ubiquitin-specific protease 2b negatively regulates IFN-β production and antiviral activity by targeting TANK-binding kinase 1. J. Immunol. 2014;193:2230–2237.
    1. McWhirter SM, et al. IFN-regulatory factor 3-dependent gene expression is defective in Tbk1-deficient mouse embryonic fibroblasts. Proc. Natl Acad. Sci. USA. 2004;101:233–238.
    1. Bulek K, et al. The inducible kinase IKKi is required for IL-17-dependent signaling associated with neutrophilia and pulmonary inflammation. Nat. Immunol. 2011;12:844–852.
    1. Hemmi H, et al. The roles of two IκB kinase-related kinases in lipopolysaccharide and double stranded RNA signaling and viral infection. J. Exp. Med. 2004;199:1641–1650.
    1. Tenoever BR, et al. Multiple functions of the IKK-related kinase IKKε in interferon-mediated antiviral immunity. Science. 2007;315:1274–1278.
    1. Hasan M, Yan N. Therapeutic potential of targeting TBK1 in autoimmune diseases and interferonopathies. Pharmacol. Res. 2016;111:336–342.
    1. Hasan M, et al. Cutting edge: inhibiting TBK1 by compound II ameliorates autoimmune disease in mice. J. Immunol. 2015;195:4573–4577.
    1. Yu J, et al. Regulation of T-cell activation and migration by the kinase TBK1 during neuroinflammation. Nat. Commun. 2015;6:6074.
    1. Zarnegar BJ, et al. Noncanonical NF-κB activation requires coordinated assembly of a regulatory complex of the adaptors cIAP1, cIAP2, TRAF2 and TRAF3 and the kinase NIK. Nat. Immunol. 2008;9:1371–1378.
    1. Willmann KL, et al. Biallelic loss-of-function mutation in NIK causes a primary immunodeficiency with multifaceted aberrant lymphoid immunity. Nat. Commun. 2014;5:5360.
    1. Li Y, et al. Cell intrinsic role of NF-κB-inducing kinase in regulating T cell-mediated immune and autoimmune responses. Sci. Rep. 2016;6:22115.
    1. Katakam AK, et al. Dendritic cells require NIK for CD40-dependent cross-priming of CD8+ T cells. Proc. Natl Acad. Sci. USA. 2015;112:14664–14669.
    1. Brightbill HD, et al. NF-κB inducing kinase is a therapeutic target for systemic lupus erythematosus. Nat. Commun. 2018;9:179.
    1. Brightbill HD, et al. Conditional deletion of NF-κB-inducing kinase (NIK) in adult mice disrupts mature B cell survival and activation. J. Immunol. 2015;195:953–964.
    1. Karnell JL, Rieder SA, Ettinger R, Kolbeck R. Targeting the CD40–CD40L pathway in autoimmune diseases: humoral immunity and beyond. Adv. Drug Deliv. Rev. 2019;141:92–103.
    1. Tsai KL, et al. A conserved mediator-CDK8 kinase module association regulates mediator–RNA polymerase II interaction. Nat. Struct. Mol. Biol. 2013;20:611–619.
    1. Malumbres M, Barbacid M. Cell cycle, CDKs and cancer: a changing paradigm. Nat. Rev. Cancer. 2009;9:153–166.
    1. Cee VJ, Chen DY, Lee MR, Nicolaou KC. Cortistatin A is a high-affinity ligand of protein kinases ROCK, CDK8, and CDK11. Angew Chem. Int. Ed. Engl. 2009;48:8952–8957.
    1. Porter DC, et al. Cyclin-dependent kinase 8 mediates chemotherapy-induced tumor-promoting paracrine activities. Proc. Natl Acad. Sci. USA. 2012;109:13799–13804.
    1. Mallinger A, et al. Discovery of potent, selective, and orally bioavailable small-molecule modulators of the mediator complex-associated kinases CDK8 and CDK19. J. Med. Chem. 2016;59:1078–1101.
    1. Czodrowski P, et al. Structure-based optimization of potent, selective, and orally bioavailable CDK8 inhibitors discovered by high-throughput screening. J. Med. Chem. 2016;59:9337–9349.
    1. Johannessen L, et al. Small-molecule studies identify CDK8 as a regulator of IL-10 in myeloid cells. Nat. Chem. Biol. 2017;13:1102–1108.
    1. Guo Z, Wang G, Lv Y, Wan YY, Zheng J. Inhibition of Cdk8/Cdk19 activity promotes Treg cell differentiation and suppresses autoimmune diseases. Front. Immunol. 2019;10:1988.
    1. Akamatsu M, et al. Conversion of antigen-specific effector/memory T cells into Foxp3-expressing Treg cells by inhibition of CDK8/19. Sci. Immunol. 2019 doi: 10.1126/sciimmunol.aaw2707.
    1. Han S, Toker A, Liu ZQ, Ohashi PS. Turning the tide against regulatory T cells. Front. Oncol. 2019;9:279.
    1. Sellwood MA, Ahmed M, Segler MH, Brown N. Artificial intelligence in drug discovery. Future Med. Chem. 2018;10:2025–2028.
    1. van Beuge MM, Poelstra K, Prakash J. Specific delivery of kinase inhibitors in nonmalignant and malignant diseases. Expert Opin. Drug Deliv. 2012;9:59–70.
    1. Cho JH, Feldman M. Heterogeneity of autoimmune diseases: pathophysiologic insights from genetics and implications for new therapies. Nat. Med. 2015;21:730–738.
    1. Uematsu S, et al. Interleukin-1 receptor-associated kinase-1 plays an essential role for Toll-like receptor (TLR)7- and TLR9-mediated interferon-α induction. J. Exp. Med. 2005;201:915–923.
    1. Vallabhapurapu S, Karin M. Regulation and function of NF-κB transcription factors in the immune system. Annu. Rev. Immunol. 2009;27:693–733.
    1. Dempsey A, Bowie AG. Innate immune recognition of DNA: a recent history. Virology. 2015;479–480:146–152.

Source: PubMed

3
Se inscrever