Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

ATP-competitive partial antagonists of the IRE1α RNase segregate outputs of the UPR

Abstract

The unfolded protein response (UPR) homeostatically matches endoplasmic reticulum (ER) protein-folding capacity to cellular secretory needs. However, under high or chronic ER stress, the UPR triggers apoptosis. This cell fate dichotomy is promoted by differential activation of the ER transmembrane kinase/endoribonuclease (RNase) IRE1α. We previously found that the RNase of IRE1α can be either fully activated or inactivated by ATP-competitive kinase inhibitors. Here we developed kinase inhibitors, partial antagonists of IRE1α RNase (PAIRs), that partially antagonize the IRE1α RNase at full occupancy. Biochemical and structural studies show that PAIRs promote partial RNase antagonism by intermediately displacing the helix αC in the IRE1α kinase domain. In insulin-producing β-cells, PAIRs permit adaptive splicing of Xbp1 mRNA while quelling destructive ER mRNA endonucleolytic decay and apoptosis. By preserving Xbp1 mRNA splicing, PAIRs allow B cells to differentiate into immunoglobulin-producing plasma cells. Thus, an intermediate RNase-inhibitory ‘sweet spot’, achieved by PAIR-bound IRE1α, captures a desirable conformation for drugging this master UPR sensor/effector.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: IRE1α dimerization is modulated by ATP-binding site occupancy.
Fig. 2: Partial antagonists of IRE1α RNase activity.
Fig. 3: Cocrystal structure of the IRE1α–PAIR1 complex.
Fig. 4: PAIR1 spares XBP1 splicing but blocks RIDD outputs in cells.
Fig. 5: PAIR1 inhibits IRE1α-dependent mRNA decay and apoptosis.
Fig. 6: PAIRs and KIRAs differentially affect B cell differentiation.

Similar content being viewed by others

Data availability

Kinome selectivity data are provided in Supplementary Dataset 1. Coordinates and structure factors have been deposited in the PDB under accession code 7BMK. Cell line RNA-seq data have been deposited to the Gene Expression Omnibus (GEO)48 and are accessible through GEO Series accession number GSE164496 (https://www.ncbi.nlm.nih.gov/geo/query/acc.cgi?acc=GSE164496). The KEGG database used in this study was acquired from http://www.genome.jp/kegg/. Source data are provided with this paper.

References

  1. Walter, P. & Ron, D. The unfolded protein response: from stress pathway to homeostatic regulation. Science 334, 1081–1086 (2011).

    Article  CAS  PubMed  Google Scholar 

  2. Hetz, C. & Papa, F. R. The unfolded protein response and cell fate control. Mol. Cell 69, 169–181 (2018).

    Article  CAS  PubMed  Google Scholar 

  3. Cox, J. S., Shamu, C. E. & Walter, P. Transcriptional induction of genes encoding endoplasmic reticulum resident proteins requires a transmembrane protein kinase. Cell 73, 1197–1206 (1993).

    Article  CAS  PubMed  Google Scholar 

  4. Mori, K., Ma, W., Gething, M.-J. & Sambrook, J. A transmembrane protein with a cdc2+/CDC28-related kinase activity is required for signaling from the ER to the nucleus. Cell 74, 743–756 (1993).

    Article  CAS  PubMed  Google Scholar 

  5. Oakes, S. A. & Papa, F. R. The role of endoplasmic reticulum stress in human pathology. Annu. Rev. Pathol. 10, 173–194 (2015).

    Article  CAS  PubMed  Google Scholar 

  6. Han, D. et al. IRE1ɑ kinase activation modes control alternate endoribonuclease outputs to determine divergent cell fates. Cell 138, 562–575 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Yoshida, H., Matsui, T., Yamamoto, A., Okada, T. & Mori, K. XBP1 mRNA is induced by ATF6 and spliced by IRE1 in response to ER stress to produce a highly active transcription factor. Cell 107, 881–891 (2001).

    Article  CAS  PubMed  Google Scholar 

  8. Calfon, M. et al. IRE1 couples endoplasmic reticulum load to secretory capacity by processing the XBP-1 mRNA. Nature 415, 92–96 (2002).

    Article  CAS  PubMed  Google Scholar 

  9. Tirasophon, W., Welihinda, A. A. & Kaufman, R. J. A stress response pathway from the endoplasmic reticulum to the nucleus requires a novel bifunctional protein kinase/endoribonuclease (Ire1p) in mammalian cells. Genes Dev. 12, 1812–1824 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Ghosh, R. et al. Allosteric inhibition of the IRE1ɑ RNase preserves cell viability and function during endoplasmic reticulum stress. Cell 158, 534–548 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Hollien, J. & Weissman, J. S. Decay of endoplasmic reticulum-localized mRNAs during the unfolded protein response. Science 313, 104–107 (2006).

    Article  CAS  PubMed  Google Scholar 

  12. Hollien, J. et al. Regulated Ire1-dependent decay of messenger RNAs in mammalian cells. J. Cell Biol. 186, 323–331 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Wang, L. et al. Divergent allosteric control of the IRE1ɑ endoribonuclease using kinase inhibitors. Nat. Chem. Biol. 8, 982–989 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Feldman, H. C. et al. Structural and functional analysis of the allosteric inhibition of IRE1ɑ with ATP-competitive ligands. ACS Chem. Biol. 11, 2195–2205 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Mendez, A. S. et al. Endoplasmic reticulum stress-independent activation of unfolded protein response kinases by a small molecule ATP-mimic. eLife 4, e05434 (2015).

    Article  PubMed Central  Google Scholar 

  16. Ferri, E. et al. Activation of the IRE1 RNase through remodeling of the kinase front pocket by ATP-competitive ligands. Nat. Commun. 11, 6387 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Lee, K. P. et al. Structure of the dual enzyme Ire1 reveals the basis for catalysis and regulation in nonconventional RNA splicing. Cell 132, 89–100 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Korennykh, A. V. et al. Cofactor-mediated conformational control in the bifunctional kinase/RNase Ire1. BMC Biol. 9, 48 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Prischi, F., Nowak, P. R., Carrara, M. & Ali, M. M. Phosphoregulation of Ire1 RNase splicing activity. Nat. Commun. 5, 3554 (2014).

    Article  PubMed  CAS  Google Scholar 

  20. Maly, D. J. & Papa, F. R. Druggable sensors of the unfolded protein response. Nat. Chem. Biol. 10, 892–901 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Harrington, P. E. et al. Unfolded protein response in cancer: IRE1ɑ inhibition by selective kinase ligands does not impair tumor cell viability. ACS Med. Chem. Lett. 6, 68–72 (2015).

    Article  CAS  PubMed  Google Scholar 

  22. Sanches, M. et al. Structure and mechanism of action of the hydroxy–aryl–aldehyde class of IRE1 endoribonuclease inhibitors. Nat. Commun. 5, 4202 (2014).

    Article  CAS  PubMed  Google Scholar 

  23. Bantscheff, M. et al. Quantitative chemical proteomics reveals mechanisms of action of clinical ABL kinase inhibitors. Nat. Biotechnol. 25, 1035–1044 (2007).

    Article  CAS  PubMed  Google Scholar 

  24. Golkowski, M. et al. Kinobead and single-shot LC–MS profiling identifies selective PKD inhibitors. J. Proteome Res. 16, 1216–1227 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Golkowski, M. et al. Kinome chemoproteomics characterization of pyrrolo[3,4-c]pyrazoles as potent and selective inhibitors of glycogen synthase kinase 3. Mol. Omics 14, 26–36 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Jan, C. H., Williams, C. C. & Weissman, J. S. Principles of ER cotranslational translocation revealed by proximity-specific ribosome profiling. Science 346, 1257521 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  27. Plumb, R., Zhang, Z. R., Appathurai, S. & Mariappan, M. A functional link between the co-translational protein translocation pathway and the UPR. eLife 4, e07426 (2015).

    Article  PubMed Central  Google Scholar 

  28. Lerner, A. G. et al. IRE1ɑ induces thioredoxin-interacting protein to activate the NLRP3 inflammasome and promote programmed cell death under irremediable ER stress. Cell Metab. 16, 250–264 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Reimold, A. M. et al. Plasma cell differentiation requires the transcription factor XBP-1. Nature 412, 300–307 (2001).

    Article  CAS  PubMed  Google Scholar 

  30. Iwakoshi, N. N. et al. Plasma cell differentiation and the unfolded protein response intersect at the transcription factor XBP-1. Nat. Immunol. 4, 321–329 (2003).

    Article  CAS  PubMed  Google Scholar 

  31. Shaffer, A. L. et al. XBP1, downstream of Blimp-1, expands the secretory apparatus and other organelles, and increases protein synthesis in plasma cell differentiation. Immunity 21, 81–93 (2004).

    Article  CAS  PubMed  Google Scholar 

  32. Tirosh, B., Iwakoshi, N. N., Glimcher, L. H. & Ploegh, H. L. XBP-1 specifically promotes IgM synthesis and secretion, but is dispensable for degradation of glycoproteins in primary B cells. J. Exp. Med. 202, 505–516 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Kawahara, T., Yanagi, H., Yura, T. & Mori, K. Endoplasmic reticulum stress-induced mRNA splicing permits synthesis of transcription factor Hac1p/Ern4p that activates the unfolded protein response. Mol. Biol. Cell 8, 1845–1862 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Cox, J. S. & Walter, P. A novel mechanism for regulating activity of a transcription factor that controls the unfolded protein response. Cell 87, 391–404 (1996).

    Article  CAS  PubMed  Google Scholar 

  35. Kimmig, P. et al. The unfolded protein response in fission yeast modulates stability of select mRNAs to maintain protein homeostasis. eLife 1, e00048 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  36. Papa, F. R., Zhang, C., Shokat, K. & Walter, P. Bypassing a kinase activity with an ATP-competitive drug. Science 302, 1533–1537 (2003).

    Article  CAS  PubMed  Google Scholar 

  37. Tong, M. & Seeliger, M. A. Targeting conformational plasticity of protein kinases. ACS Chem. Biol. 10, 190–200 (2015).

    Article  CAS  PubMed  Google Scholar 

  38. Jura, N. et al. Catalytic control in the EGF receptor and its connection to general kinase regulatory mechanisms. Mol. Cell 42, 9–22 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Zhang, K. et al. The unfolded protein response sensor IRE1ɑ is required at 2 distinct steps in B cell lymphopoiesis. J. Clin. Invest. 115, 268–281 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Iwawaki, T., Akai, R., Yamanaka, S. & Kohno, K. Function of IRE1ɑ in the placenta is essential for placental development and embryonic viability. Proc. Natl Acad. Sci. USA 106, 16657–16662 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. S. Morita, S. A. et al. Targeting ABL–IRE1α signaling spares ER-stressed pancreatic β-cells to reverse autoimmune diabetes. Cell Metab. 25, 883–897 (2017).

    Article  CAS  Google Scholar 

  42. Sato, H., Shiba, Y., Tsuchiya, Y., Saito, M. & Kohno, K. 4μ8C inhibits insulin secretion independent of IRE1α RNase activity. Cell Struct. Funct. 42, 61–70 (2017).

    Article  CAS  PubMed  Google Scholar 

  43. Papandreou, I. et al. Identification of an Ire1ɑ endonuclease specific inhibitor with cytotoxic activity against human multiple myeloma. Blood 117, 1311–1314 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Grandjean, J. et al. Pharmacologic IRE1/XBP1s activation confers targeted ER proteostasis reprogramming. Nat. Chem. Biol. 16, 1052–1061 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Yu, Y., Pieper, R. & Smith, M. A spinnable and automatable StageTip for high throughput peptide desalting and proteomics. Protoc. Exch. https://doi.org/10.1038/protex.2014.033 (2014).

  46. Metz, K. S. et al. Coral: clear and customizable visualization of human kinome data. Cell Syst. 7, 347–350 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Qi, L., Yang, L. & Chen, H. Detecting and quantitating physiological endoplasmic reticulum stress. Methods Enzymol. 490, 137–146 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Edgar, R., Domrachev, M. & Lash, A. E. Gene Expression Omnibus: NCBI gene expression and hybridization array data repository. Nucleic Acids Res. 30, 207–210 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This work was supported by the Leona M. and Harry B. Helmsley Charitable Trust (G-3231 to F.R.P.), the Juvenile Diabetes Research Foundation (JDRF 2-SRA-2016-234-S-N to F.R.P.) and the NIH (R01DK116064 to D.J.M. and F.R.P.; R01DK100623 to B.J.B, D.J.M. and F.R.P.; U01DK123609 to D.J.M. and F.R.P.; R01HL145037 to F.R.P.; K08HL157654 to V.C.A.; R01AR069520 to J.Z. and R01AI148487 to J.Z.).

Author information

Authors and Affiliations

Authors

Contributions

H.C.F. executed all enzymatic and biochemical analyses. H.C.F. and B.J.B. designed compounds. H.C.F. and V.N.V. synthesized and characterized compounds. R.G., J-H.K. and A.O. performed cellular analysis of XBP1 splicing, RIDD and apoptosis. V.C.A. conducted RNA-seq analyses. H.C.F. and Z.E.P. performed target engagement studies. H.C.F. and B.G.K.P. performed chemoproteomic profiling. Plasma cell differentiation studies were designed by J.Z. and performed by J.L.M. J.Z., F.R.P. and D.J.M. supervised experiments. H.C.F., F.R.P. and D.J.M. wrote the manuscript with input from all authors. F.R.P. and D.J.M. conceived of and led this study.

Corresponding authors

Correspondence to Feroz R. Papa or Dustin J. Maly.

Ethics declarations

Competing interests

B.J.B., F.R.P. and D.J.M. are founders, equity holders and consultants for OptiKIRA, LLC (Cleveland, OH), a biotech company founded on the treatment of ER-stress-induced retinal degeneration; no company funding or chemical matter supported the work in this manuscript. J.Z. serves on the SAB for Walking Fish Therapeutics.

Additional information

Peer review information Nature Chemical Biology thanks Nathanael Gray, Boaz Tirosh and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 KIRAs monomerize IRE1α*.

(a) Chemical structure of compound 2. (b) Inhibition of IRE1α*’s kinase activity by compound 2 (Ki = 480 nM). Data points shown are the mean of n = 3 independent experiments ± SEM. (c) Kdimer curve of the IRE1α*-compound 2 complex (teal). Kdimer curves of apo IRE1α* (dark gray) and the IRE1α*-KIRA8 complex (light blue) from Fig. 1e are shown for comparison. Data points shown are the mean of n = 3 independent experiments ± SEM.

Extended Data Fig. 2 KIRA8 completely suppresses XBP1 splicing in ER-stressed INS-1 cells.

(a) Quantification of spliced XBP1 in untreated parent INS-1 cells and parent INS-1 cells treated with brefeldin A (BFA) or BFA and KIRA8. Values shown are the mean of n = 3 biologically independent samples ± SEM. (b) Representative example of an EtBr-stained agarose gel of XBP1 cDNA amplicons from INS-1 cells subjected to the conditions described in (a). (c) Quantification of spliced XBP1 in untreated parent INS-1 cells and parent INS-1 cells treated with thapsigargin (Tg) or Tg and KIRA8. Values shown are the mean of n = 3 biologically independent samples ± SEM. (d) Representative example of an EtBr-stained agarose gel of XBP1 cDNA amplicons from INS-1 cells subjected to the conditions described in (c).

Source data

Extended Data Fig. 3 Structurally-related PAIRs and KIRAs.

(a) Chemical structure of (left) and inhibition of IRE1α*’s kinase activity (right, Ki = 140 nM) by compound 6. (b) Chemical structure of (left) and inhibition of IRE1α*’s kinase activity (right, Ki = 98 nM) by compound 7. (c) Chemical structure of (left) and inhibition of IRE1α*’s kinase activity (right, Ki = 220 nM) by compound 8 (PAIR1). (d) Chemical structure of (left) and inhibition of IRE1α*’s kinase activity (right, Ki = 160 nM) by compound 9 (KIRA9). Inhibition of IRE1α*’s RNase activity by (e) compounds 4 and 5, and (f) PAIR1 and KIRA9. All data points shown are the mean of n = 3 independent experiments ± SEM.

Extended Data Fig. 4 Partial antagonism of RNase activity is directly related to R3-group size.

Three-way correlation of kinase inhibition (a proxy for ATP-binding site occupancy), residual RNase activity, and R3-group size measured as Connolly molecular surface area (Å2) from matched inhibitor sets: (a) 4 and 5, (b) 10 (a previously described KIRA), 6, 7, and (c) KIRA9 and PAIR1. Kinase inhibition and residual RNase activity are shown as the mean of n = 3 independent experiments ± SEM. Note: values are also shown in Fig. 2f.

Extended Data Fig. 5 PAIR1 stabilizes an intermediately displaced helix-αC.

(a) Superimposition of the helix-αCs of active IRE1α* (gray, PDB: 5HGI) with the PAIR1-IRE1α* complex (orange), and the KIRA8-IRE1α* complex (blue, PDB: 6URC). Inset displays the displacement of Glu612 within the helix-αC relative to active IRE1α* (gray, PDB: 5HGI). (b) 2Fo-Fc electron density maps of Glu612 from active IRE1α* (gray, PDB: 5HGI), PAIR1-IRE1α* (orange), and AMG-16-IRE1α* (teal, PDB: 4U6R), contoured to 1.0σ shown as gray isomesh, revealing strong electron density for Glu612, making it a suitable residue to measure approximate helix-αC movement. (c) Displacement of β-carbons from helix-αC residues (606-617) from co-crystal structures of IRE1α* with PAIR1 (orange), AMG-16 (teal, PDB: 4U6R), and KIRA8 (blue, PDB: 6URC) relative to active IRE1α* (gray, PDB: 5HGI).

Extended Data Fig. 6 The PAIR1-bound IRE1α* dimer interface is partially disrupted through Asp 620 displacement.

(a) Superimposition of the active IRE1α* (gray, PDB: 5HGI) and PAIR1-IRE1α* complex (orange) dimers. (b,c) Zoom-in views of the dimer interface contacts that are similar between active IRE1α* (gray, PDB: 5HGI) and the PAIR1-IRE1α* complex (orange) dimers. (d) 2Fo-Fc electron density maps of Asp620, contoured to 1.0σ shown as gray isomesh, reveal equal distribution of Asp620 between ‘in’ and ‘out’ conformations. (e) Superimposition of dimer interface residue Asp620 from active IRE1α* (gray, PDB: 5HGI) and the PAIR1-IRE1α* complex (orange) shows that Asp620-IN is in a similar conformation as active IRE1α* and forms a salt bridge with the side-chains of Arg594 and Arg627 of the adjacent IRE1α* protomer. In the other conformation, Asp620-OUT, the side-chain of Asp620 is displaced 11.6 Å and can no longer form a salt bridge with Arg594 and Arg627. This inter-dimer salt bridge between Asp620 and Arg594/Arg627 is essential for RNase active dimer formation and although Asp620 is found equally in conformations productive (Asp620-IN) and unproductive (Asp620-OUT) for RNase active dimer formation, the cumulative effect is partial disruption of the RNase active dimer interface. (f) Superimposition of kinase catalytic residues: K599, Glu612, and the DFG-motif (Asp712, Gly713, Phe714) from the PAIR1-IRE1α* complex (orange) and the AMG-16-IRE1α* complex (teal, PDB:4U6R) reveal that residues outside of the helix-αC (K599 and the DFG-motif) are in similar conformations, suggesting that differences in the pharmacology between PAIRs and KIRAs stems mainly from helix-αC movement.

Extended Data Fig. 7 PAIR1 and KIRA9 equipotently block ER stress-induced IRE1α autophosphorylation.

Immunoblots for total IRE1α (bottom) from parent INS-1 cells treated with various concentrations of PAIR1 or KIRA9 followed by the addition of 200 ng/μL Brefeldin-A (BFA) for 2 hours. Samples were subjected to phostag SDS-PAGE and % non-phosphorylated IRE1α (top) was determined from the ratio of dephosphorylated IRE1α relative to total IRE1α. Values shown are the mean of n = 3 biologically independent samples ± SEM.

Source data

Extended Data Fig. 8 Characterization of the highly potent PAIR, PAIR2.

(a) Chemical structure of PAIR2. (b) Inhibition of IRE1α*’s kinase activity by PAIR2 (Ki = 8.8 nM, values shown as the mean of n = 3 independent experiments ± SEM) and KIRA8 (Ki = 12 nM, values shown as the mean of n = 2 independent experiments). (c) Immunoblots for total IRE1α (bottom) from the INS-1 cells described in Fig. 4b treated with various concentrations of PAIR2 or KIRA8 followed by the addition of 5 ng/ml doxycycline (Dox) for 6 hours. Samples were subjected to phostag SDS-PAGE and % phosphorylated IRE1α (top) was determined from the ratio of phosphorylated IRE1α relative to total IRE1α. Values shown are the mean of n = 3 independent experiments ± SEM. (d) Percent spliced XBP1 (quantified from PstI-digested XBP1 cDNA amplicons) from parent INS-1 cells treated with DMSO, PAIR2, or KIRA8 (0.5, 1, 3, 10 μM) followed by the addition of 2 mM 1,4-dithiothreitol (DTT) (top). Data shown are the mean of n = 3 biologically independent samples ± SEM. A representative EtBr-stained agarose gel of XBP1 cDNA amplicons is shown below. (e) Percent spliced XBP1 (quantified from PstI-digested XBP1 cDNA amplicons) from parent INS-1 cells treated first with DMSO (lanes 1-4) or 10 μM PAIR2 (lanes 5-7) for 30 minutes followed by treatment with either DMSO (lanes 1, 2, and 5) or 0.5 or 1 μM KIRA8 (lanes 3, 4, 6, and 7) for 30 minutes followed by the addition of 2 mM 1,4-dithiothreitol (DTT) (top). Data shown are the mean of n = 3 biologically independent samples ± SEM. A representative EtBr-stained agarose gel of XBP1 cDNA amplicons is shown below. (f) Kinome selectivity of PAIR2 as determined by kinobead profiling. Kinases that were identified in the profiling experiment are shown with gray circles. Circle size and color have been scaled to the log2 ratio (difference in LFQ intensity) between DMSO and treatment with 10 and 2.5 μM of PAIR2 (mean of three and four replicates, respectively). Kinases reported as being drug-sensitive (Log2 Difference > 2) were also required to show significance from a two-sample T-test with FDR of 0.05. (g) Heat map displaying kinome selectivity from (f). P values calculated in (d), (e) are versus DTT treated cells and were determined using a two-tailed Student’s t-test. *P < 0.05, **P < 0.01, ***P < 0.001.

Source data

Extended Data Fig. 9 Flow cytometric analysis of lipopolysaccharide (LPS)-treated mouse splenocytes.

Mouse splenocytes were loaded with vital dye and subsequently cultured +/− 1 μg/mL LPS for three days in the presence or absence of KIRA8, KIRA9, PAIR1, or PAIR2 at noted concentrations. Following culture, cells were stained to detect surface marker expression and subsequently permeabilized to detect XBP1s intracellular expression. Samples were mixed with a fixed number of counting beads and collected by flow cytometry. Representative plots are displayed. (a) Splenocytes and beads were first gated on the basis of forward scatter (FSC) and side scatter (SSC) for the purpose of absolute quantification of cell number. (b) Live B-cells were then identified on the basis of CD19 expression and exclusion of live/dead dye. (c) Plasmablasts were identified as CD19+ cells with downregulation of B220 and upregulation of CD138 expression. (d) Plots depict vital dye dilution as a marker of cell division and XBP1s + gate. Corresponding quantification of these cell populations is shown in Fig. 6 and Extended Data Fig. 10.

Extended Data Fig. 10 Effect of PAIRs and KIRAs on B-cell division.

Quantification of CD19 positive cells following treatment of mouse splenocytes with 1 μg/mL LPS for three days in the presence or absence of various concentrations of (a) PAIR1 (orange) or KIRA9 (teal) or (b) PAIR2 (dark orange) or KIRA8 (dark blue) as gated in Extended Data Fig. 9. (c, d) Division index of CD19 + cells treated as in (a, b) was calculated via Flowjo on the basis of vital dye dilution (as shown in Extended Data Fig. 9). P values were calculated using one-tailed Student’s t-test. *P < 0.033, **P < 0.002, ***P < 0.001. Data for 20 μM KIRA8 could not be generated due to cellular toxicity. All data shown are the mean of n = 3 biologically independent samples ± SEM.

Supplementary information

Supplementary Information

Supplemental Figs. 1–18, Table 1 and Note (with Datasets and Chemistry Methods).

Reporting Summary

Supplementary Data

Complete proteomic dataset for kinome selectivity profiling using HEK293T and HCT116 cells treated with inhibitors 2, 3, 6, 7, PAIR1, KIRA9 or PAIR2. Relevant for Fig. 4a, Supplemental Fig. 9 and Extended Data Fig. 8.

Source data

Source Data Fig. 4

Unprocessed westerns and agarose gels for Fig. 4b,c,i.

Source Data Fig. 5

Unprocessed westerns for Fig. 5b.

Source Data Extended Data Fig. 2

Unprocessed agarose gels for Extended Data Fig. 2.

Source Data Extended Data Fig. 7

Unprocessed westerns for Extended Data Fig. 7.

Source Data Extended Data Fig. 8

Unprocessed westerns and agarose gels for Extended Data Fig. 8c,d,e.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Feldman, H.C., Ghosh, R., Auyeung, V.C. et al. ATP-competitive partial antagonists of the IRE1α RNase segregate outputs of the UPR. Nat Chem Biol 17, 1148–1156 (2021). https://doi.org/10.1038/s41589-021-00852-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41589-021-00852-0

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing