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:

Molecular basis for pseudokinase-dependent autoinhibition of JAK2 tyrosine kinase

This article has been updated

Abstract

Janus kinase-2 (JAK2) mediates signaling by various cytokines, including erythropoietin and growth hormone. JAK2 possesses tandem pseudokinase and tyrosine-kinase domains. Mutations in the pseudokinase domain are causally linked to myeloproliferative neoplasms (MPNs) in humans. The structure of the JAK2 tandem kinase domains is unknown, and therefore the molecular bases for pseudokinase-mediated autoinhibition and pathogenic activation remain obscure. Using molecular dynamics simulations of protein-protein docking, we produced a structural model for the autoinhibitory interaction between the JAK2 pseudokinase and kinase domains. A striking feature of our model, which is supported by mutagenesis experiments, is that nearly all of the disease mutations map to the domain interface. The simulations indicate that the kinase domain is stabilized in an inactive state by the pseudokinase domain, and they offer a molecular rationale for the hyperactivity of V617F, the predominant JAK2 MPN mutation.

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

Figure 1: Steps of JAK2 JH2-JH1 model generation.
Figure 2: Model of JAK2 JH2-JH1 derived from MD simulations.
Figure 3: Experimental validation of the JAK2 JH2-JH1 model.
Figure 4: JH2-mediated autoinhibition of JH1 in JAK2.
Figure 5: Model for JAK2 JH2-mediated autoinhibition of JH1.

Similar content being viewed by others

Accession codes

Accessions

Protein Data Bank

Change history

  • 12 June 2014

    In the version of this article originally posted online, the model coordinates presented in Supplementary Data Sets 1 and 2 were switched. These errors have been corrected as of 12 June 2014.

References

  1. O'Shea, J.J., Holland, S.M. & Staudt, L.M. JAKs and STATs in immunity, immunodeficiency, and cancer. N. Engl. J. Med. 368, 161–170 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Wallweber, H.J., Tam, C., Franke, Y., Starovasnik, M.A. & Lupardus, P.J. Structural basis of recognition of interferon-α receptor by tyrosine kinase 2. Nat. Struct. Mol. Biol. 21, 443–448 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Vainchenker, W., Delhommeau, F., Constantinescu, S.N. & Bernard, O.A. New mutations and pathogenesis of myeloproliferative neoplasms. Blood 118, 1723–1735 (2011).

    Article  CAS  PubMed  Google Scholar 

  4. Haan, C., Behrmann, I. & Haan, S. Perspectives for the use of structural information and chemical genetics to develop inhibitors of Janus kinases. J. Cell. Mol. Med. 14, 504–527 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Kralovics, R. et al. A gain-of-function mutation of JAK2 in myeloproliferative disorders. N. Engl. J. Med. 352, 1779–1790 (2005).

    Article  CAS  PubMed  Google Scholar 

  6. Baxter, E.J. et al. Acquired mutation of the tyrosine kinase JAK2 in human myeloproliferative disorders. Lancet 365, 1054–1061 (2005).

    Article  CAS  PubMed  Google Scholar 

  7. Levine, R.L. et al. Activating mutation in the tyrosine kinase JAK2 in polycythemia vera, essential thrombocythemia, and myeloid metaplasia with myelofibrosis. Cancer Cell 7, 387–397 (2005).

    Article  CAS  PubMed  Google Scholar 

  8. Lipson, D. et al. Identification of new ALK and RET gene fusions from colorectal and lung cancer biopsies. Nat. Med. 18, 382–384 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Saharinen, P. & Silvennoinen, O. The pseudokinase domain is required for suppression of basal activity of Jak2 and Jak3 tyrosine kinases and for cytokine-inducible activation of signal transduction. J. Biol. Chem. 277, 47954–47963 (2002).

    Article  CAS  PubMed  Google Scholar 

  10. Saharinen, P., Takaluoma, K. & Silvennoinen, O. Regulation of the Jak2 tyrosine kinase by its pseudokinase domain. Mol. Cell. Biol. 20, 3387–3395 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Ungureanu, D. et al. The pseudokinase domain of JAK2 is a dual-specificity protein kinase that negatively regulates cytokine signaling. Nat. Struct. Mol. Biol. 18, 971–976 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Bandaranayake, R.M. et al. Crystal structures of the JAK2 pseudokinase domain and the pathogenic mutant V617F. Nat. Struct. Mol. Biol. 19, 754–759 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Shan, Y. et al. How does a drug molecule find its target binding site? J. Am. Chem. Soc. 133, 9181–9183 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Dusa, A., Mouton, C., Pecquet, C., Herman, M. & Constantinescu, S.N. JAK2 V617F constitutive activation requires JH2 residue F595: a pseudokinase domain target for specific inhibitors. PLoS ONE 5, e11157 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  15. Liang, S., Liu, S., Zhang, C. & Zhou, Y. A simple reference state makes a significant improvement in near-native selections from structurally refined docking decoys. Proteins 69, 244–253 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Liang, S., Zhang, C., Sarmiento, J. & Standley, D.M. Protein loop modeling with optimized backbone potential functions. J. Chem. Theory Comput. 8, 1820–1827 (2012).

    Article  CAS  PubMed  Google Scholar 

  17. Argetsinger, L.S. et al. Autophosphorylation of JAK2 on tyrosines 221 and 570 regulates its activity. Mol. Cell. Biol. 24, 4955–4967 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Feener, E.P., Rosario, F., Dunn, S.L., Stancheva, Z. & Myers, M.G. Tyrosine phosphorylation of Jak2 in the JH2 domain inhibits cytokine signaling. Mol. Cell. Biol. 24, 4968–4978 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Zhao, L. et al. A JAK2 interdomain linker relays Epo receptor engagement signals to kinase activation. J. Biol. Chem. 284, 26988–26998 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Ishida-Takahashi, R. et al. Phosphorylation of Jak2 on Ser523 inhibits Jak2-dependent leptin receptor signaling. Mol. Cell. Biol. 26, 4063–4073 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Mazurkiewicz-Munoz, A.M. et al. Phosphorylation of JAK2 at serine 523: a negative regulator of JAK2 that is stimulated by growth hormone and epidermal growth factor. Mol. Cell. Biol. 26, 4052–4062 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Mullighan, C.G. et al. JAK mutations in high-risk childhood acute lymphoblastic leukemia. Proc. Natl. Acad. Sci. USA 106, 9414–9418 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Wan, X., Ma, Y., McClendon, C.L., Huang, L.J. & Huang, N. Ab initio modeling and experimental assessment of Janus Kinase 2 (JAK2) kinase-pseudokinase complex structure. PLoS Comput. Biol. 9, e1003022 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Miyazawa, S. & Jernigan, R.L. Residue-residue potentials with a favorable contact pair term and an unfavorable high packing density term, for simulation and threading. J. Mol. Biol. 256, 623–644 (1996).

    Article  CAS  PubMed  Google Scholar 

  25. Nagar, B. et al. Structural basis for the autoinhibition of c-Abl tyrosine kinase. Cell 112, 859–871 (2003).

    Article  CAS  PubMed  Google Scholar 

  26. Huse, M. & Kuriyan, J. The conformational plasticity of protein kinases. Cell 109, 275–282 (2002).

    Article  CAS  PubMed  Google Scholar 

  27. Shan, Y. et al. A conserved protonation-dependent switch controls drug binding in the Abl kinase. Proc. Natl. Acad. Sci. USA 106, 139–144 (2009).

    Article  CAS  PubMed  Google Scholar 

  28. Gnanasambandan, K., Magis, A. & Sayeski, P.P. The constitutive activation of Jak2–V617F is mediated by a π stacking mechanism involving phenylalanines 595 and 617. Biochemistry 49, 9972–9984 (2010).

    Article  CAS  PubMed  Google Scholar 

  29. Lindauer, K., Loerting, T., Liedl, K.R. & Kroemer, R.T. Prediction of the structure of human Janus kinase 2 (JAK2) comprising the two carboxy-terminal domains reveals a mechanism for autoregulation. Protein Eng. 14, 27–37 (2001).

    Article  CAS  PubMed  Google Scholar 

  30. Wan, S. & Coveney, P.V. Regulation of JAK2 activation by Janus homology 2: evidence from molecular dynamics simulations. J. Chem. Inf. Model. 52, 2992–3000 (2012).

    Article  CAS  PubMed  Google Scholar 

  31. Toms, A.V. et al. Structure of a pseudokinase-domain switch that controls oncogenic activation of Jak kinases. Nat. Struct. Mol. Biol. 20, 1221–1223 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Lupardus, P.J., et al. Structure of the pseudokinase-kinase domains from protein kinase TYK2 reveals a mechanism for Janus kinase (JAK) autoinhibition. Proc. Natl. Acad. Sci. USA 111, 8025–8030 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  33. LaFave, L.M. & Levine, R.L. JAK2 the future: therapeutic strategies for JAK-dependent malignancies. Trends Pharmacol. Sci. 33, 574–582 (2012).

    Article  CAS  PubMed  Google Scholar 

  34. Baffert, F. et al. Potent and selective inhibition of polycythemia by the quinoxaline JAK2 inhibitor NVP-BSK805. Mol. Cancer Ther. 9, 1945–1955 (2010).

    Article  CAS  PubMed  Google Scholar 

  35. Best, R.B. et al. Optimization of the additive CHARMM all-atom protein force field targeting improved sampling of the backbone Φ, Ψ and side-chain χ1 and χ2 dihedral angles. J. Chem. Theory Comput. 8, 3257–3273 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Mackerell, A.D. Jr. et al. All-atom empirical potential for molecular modeling and dynamics studies of proteins. J. Phys. Chem. B 102, 3586–3616 (1998).

    Article  CAS  PubMed  Google Scholar 

  37. Jorgensen, W.L., Chandrasekhar, J., Madura, J.D., Impey, R.W. & Klein, M.L. Comparison of simple potential functions for simulating liquid water. J. Chem. Phys. 79, 926–935 (1983).

    Article  CAS  Google Scholar 

  38. Shaw, D.E. et al. in Proc. of the ACAM/IEEE Conference on Supercomputing Vol. 51, 91–97 (ACM Press, New York, 2008).

    Google Scholar 

  39. Hoover, W.G. Canonical dynamics: equilibrium phase-space distributions. Phys. Rev. A 31, 1695–1697 (1985).

    Article  CAS  Google Scholar 

  40. Lippert, R.A. et al. A common, avoidable source of error in molecular dynamics integrators. J. Chem. Phys. 126, 046101 (2007).

    Article  PubMed  Google Scholar 

  41. Krautler, V., Van Gunsteren, W.F. & Hunenberger, P.H. A fast SHAKE algorithm to solve distance constraint equations for small molecules in molecular dynamics simulations. J. Comput. Chem. 22, 501–508 (2001).

    Article  CAS  Google Scholar 

  42. Fennell, C.J. & Gezelter, J.D. Is the Ewald summation still necessary? Pairwise alternatives to the accepted standard for long-range electrostatics. J. Chem. Phys. 124, 234104 (2006).

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

This work was supported in part by the US National Institutes of Health (grant R21 AI095808; S.R.H.) and the Medical Research Council of Academy of Finland, Sigrid Juselius Foundation, Medical Research Fund of Tampere University Hospital, Finnish Cancer Foundation and Tampere Tuberculosis Foundation (O.S.). The APRE-Rluc and pRG-TK plasmids were gifts from D. Levy and J. Belasco (both at New York University School of Medicine), respectively. We thank R. Bandaranayake for crystallographic support, A. Philippsen for animation support and W.T. Miller, M. Mohammadi and M.P. Eastwood for critical reading of the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

Y.S. conceived the project, supervised MD simulations and wrote the manuscript; K.G. performed biochemical experiments and wrote the manuscript; D.U. and H.H. performed biochemical experiments and protein production; E.T.K. performed MD simulations; K.Y. performed protein docking analysis; O.S. supervised biochemical studies and wrote the manuscript; D.E.S. supervised MD simulations and wrote the manuscript; and S.R.H. conceived the project, supervised biochemical experiments and wrote the manuscript.

Corresponding authors

Correspondence to Yibing Shan, David E Shaw or Stevan R Hubbard.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Analysis of the initial 14 JAK2 JH2-JH1 simulations, and contact surfaces in the final model.

(a) Analysis of 14 JAK2 JH2-JH1 simulations. From each of the 14 3.0-µs simulations, starting from an arbitrary JH2-JH1 non-contacting pose, 300 snapshots (10-ns interval) were evaluated using both EMPIRE and OSCAR scoring functions15,16. The score of each snapshot from each simulation was plotted in the two-dimensional energy space with a unique color. The dots corresponding to the simulation that generated pose 2, the JH2-JH1 interaction that was pursued further (see Fig. 1), are colored red. (b) Time evolution of the orientation of JH1 and its center of mass for each of the 14 initial simulations (see Fig. 1). For each time snapshot in the simulations, the JH1 orientation is represented by the angles of the JH1 principal axes of inertia relative to the principal axes of JH1 in the final model. Also shown for each snapshot is the distance of the JH1 center-of-mass from the JH1 center-of-mass in the final model. JH2 in each snapshot is superimposed on JH2 in the final model prior to the angle and distance calculations. As shown, pose 2 (end of simulation 2) bears resemblance to, but is non-trivially different from the pose in the final model. Only after addition of the JH2-JH1 linker and further simulation (2 Extended) does this pose converge towards the final model. Other than simulation 2, simulation 12 also reached a JH2-JH1 pose that bears resemblance to the pose in the final model. (c) Surface representations of JH2 and JH1 (from final model) in “open book” view, in which JH2 has been rotated clockwise by 90° (vertical axis) and JH1 counterclockwise by 90°, with respect to the orientation in Fig. 2a, to reveal the interaction surface. Left, in cyan outline are residues in JH2 (residues 537–810) that are within 4.0 Å of an atom in JH1, and in green outline are residues in JH2 that are within 4.0 Å of an atom in the SH2-JH2 linker (residues 520–536). Right, same as left, except for JH1 (residues 840–1132) relative to JH2 (orange outline) and the SH2-JH2 linker (green outline). Activating mutations in JH1 and JH2 are colored pink and labeled. Other residues of interest are also labeled, as are the N and C lobes of the kinase domains.

Supplementary Figure 2 Functional studies of JAK2 mutants.

(a) Luciferase activities of wild-type and mutant JAK2 measured using an APRE-luc reporter to assess endogenous STAT3-dependent transcription in COS7 cells. The firefly luciferase activity of each sample was normalized to that of renilla luciferase (luciferase ratio) and plotted as fold-change relative to the wild-type JAK2 (WT) luciferase ratio (set to 1.0). Average values and standard deviations were derived from triplicate samples. (b) Analysis of JAK2 mutants D873R and D873N in JH1 alone. Representative western blot of JAK2 JH1 (HA-tagged) immunoprecipitated from transfected HEK 293T cells and probed for JAK2 Tyr1007-Tyr1008 phosphorylation (pJAK2) (top) or protein levels (HA) (bottom). A reference molecular-weight marker (in kDa) is indicated on the left. (c) Epo-dependent activation of JAK2 mutants. JAK2-deficient γ2A cells transfected with the indicated JAK2 and STAT5 (HA-tagged) plasmids were either left untreated (−) or stimulated with Epo (+), and the whole cell lysates were probed for JAK2 Tyr1007-Tyr1008 phosphorylation (pJAK2, top), STAT5 phosphorylation (pSTAT5, middle), or protein levels (HA, bottom). Reference molecular-weight markers (in kDa) are indicated on the left.

Supplementary Figure 3 Simulations of JAK2 JH2-JH1 mutants.

(a) Analysis of R683E, D873N and R683E D873N. The distance between the Cα atoms of Glu592 (JH2) and Arg947 (JH1), interacting residues in the JH2-JH1 model (see Fig. 2e), is plotted as a function of simulation time. In the activating mutations R683E and D873N, the two residues (Glu592, Arg947) separated, whereas for wild type (WT) and the double mutation, the distance remained relatively stable. (b) Salt-bridge analysis for JAK2 Glu592-Arg947. WT, E592R and E592R R947E were simulated for 7.5 µs each. Plotted are the distances between select residues as a function of simulation time. Shown in solid lines are the distance trajectories between native residues, and shown in dashed lines are distance trajectories in which at least one of the residues involved has been introduced by mutation. To simplify the salt-bridge presentation, the actual distances displayed are between Cζ of Arg588, Arg592 or Arg947 (to account for Nɛ, Nη1, and Nη2 of arginine) and either Cδ of Glu592 or Glu947 (to account for Oɛ1 and Oɛ2 of glutamic acid) or P of pSer523 (to account for O1P, O2P and O3P of phosphoserine). Thus, the representative distance for a salt bridge is 3.8 Å rather than 2.7 Å (typical nitrogen-oxygen distance). Gray regions indicate the approximate distance range for salt bridges.

Supplementary Figure 4 Conformational changes in JH1 and analysis of V617F.

Conformational changes in JH1 and analysis of V617F. (a) Radius of gyration (Rg, an overall measure of the size) of JH1 as a function of simulation time (same simulations as in Fig. 4a). (b) Root-mean-square deviation (RMSD) for Cα atoms in JH1 as a function of simulation time, after aligning Cα atoms in JH2 for each time frame with the Cα atoms in their initial positions. Both simulations were initiated from an identical JH2-JH1 pose with Ser523 and Tyr570 phosphorylated. In one simulation, the JH1 activation loop was phosphorylated at Tyr1007 and Tyr1008 (orange), and in the other simulation (green), the activation loop was unphosphorylated. A high RMSD is indicative of a high degree of structural deviation from the JH2-JH1 configuration (the autoinhibitory pose) shown in Fig. 2a. (ce) Analyses of JAK2 MPN mutation V617F. In previous studies, Ungureanu et al.11 showed that Ser523 and Tyr570 are poorly phosphorylated in activated JAK2 mutants such as V617F. Thus, to simulate V617F JH2-JH1 properly, we left Ser523 (and Tyr570) unphosphorylated and, for comparison purposes, we also simulated unphosphorylated wild type (WT). F595A, in αC of JH2 and proximal to Val617, was shown to suppress V617F14,28, and therefore we also simulated F595A V617F (unphosphorylated). (c) The SH2-JH2 linker conformations visited during the simulations. JH2-JH1 in the first time frame (t = 0) of each trajectory is shown in ribbon representation and colored as in Fig. 1. The Cα trace of the SH2-JH2 linker (residues 522–536) for each time frame is shown in green, after aligning JH2 in each frame with JH2 at t = 0. As shown, the linker in V617F is least stable in the binding groove between JH2 and JH1. (d) RMSD for Cα atoms in JH1 as a function of simulation time, after aligning Cα atoms in JH2 for each time frame with the Cα atoms in their initial positions. Removal of pSer523 and pTyr570 (WT versus WT pSpY) led to an increase in the overall conformational heterogeneity of JH1 and thus destabilization of the JH2-JH1 complex. V617F caused a further increase in JH1 heterogeneity, and addition of F595A to V617F restored JH1 heterogeneity back to the level of phosphorylated JH2-JH1 (WT pSpY). (e) RMSD for Cα atoms in αC of JH1 (residues 889–904) relative to the active conformation, after aligning all the Cα atoms in JH1. As shown, the active conformation of αC is most stable in V617F.

Supplementary Figure 5 JAK1 JH2-JH1 model and comparison of JAK2 JH2-JH1 models.

(a) Atomic models of JAK1 JH2 (PDB 4L00 (ref. 31)) and JH1 (PDB 4E5W (Kulagowski, J.J., et al. J. Med. Chem. 55, 5901–5921 (2012)) were placed by superposition into the positions of JH2 and JH1 of JAK2 (Fig. 2a). The SH2-JH2 and JH2-JH1 linkers were added, and an MD simulation was run for 12 µs. Shown is a representative pose near the end of the simulation. JH2 (residues 575–850) is colored orange, JH1 (residues 866–1154) is colored cyan, with the activation loop (residues 1021–1043) colored red, the SH2-JH2 linker (residues 563–574) is colored green, and the JH2-JH1 linker (residues 851–865) is colored gray. Mapped activating mutations (Hornakova, T., et al. Haematologica 96, 845–853 (2011)) are shown in stick representation, colored pink (carbon atoms), and labeled. Red, oxygen atoms; blue, nitrogen atoms. Other residues of interest are shown in stick representation and labeled. The N-terminus (residue 563) is labeled N, and the C-terminus (residue 1154) is labeled C. (b) Ribbon diagram of the JAK2 JH2-JH1 model from the current study (1), from Lindauer et al.29 (2), and from Wan et al.23 (3). Coloring is the same as in Fig. 1. The side chains of Val617, Arg683, and Asp873 are shown in sphere representation and colored pink (carbon atoms). The models are aligned with one another based on a superposition of JH2. The N- and C-termini for each model are labeled N and C.

Supplementary Figure 6 Uncropped images for western blots shown in Fig. 3.

For each blot (labeled accordingly), boxes mark the borders of the cropped images shown in Fig. 3.

Supplementary information

Supplementary Figures

Supplementary Figures 1–6 (PDF 11976 kb)

Supplementary Data Set 1

JAK2 JH2-JH1 model coordinates (TXT 396 kb)

Supplementary Data Set 2

JAK1 JH2-JH1 model coordinates (TXT 377 kb)

Supplementary Video 1

Generation of the JAK2 JH2-JH1 model (MOV 8036 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Shan, Y., Gnanasambandan, K., Ungureanu, D. et al. Molecular basis for pseudokinase-dependent autoinhibition of JAK2 tyrosine kinase. Nat Struct Mol Biol 21, 579–584 (2014). https://doi.org/10.1038/nsmb.2849

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nsmb.2849

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