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:

Targeting lipid–protein interaction to treat Syk-mediated acute myeloid leukemia

Abstract

Membrane lipids control the cellular activity of kinases containing the Src homology 2 (SH2) domain through direct lipid–SH2 domain interactions. Here we report development of new nonlipidic small molecule inhibitors of the lipid–SH2 domain interaction that block the cellular activity of their host proteins. As a pilot study, we evaluated the efficacy of lipid–SH2 domain interaction inhibitors for spleen tyrosine kinase (Syk), which is implicated in hematopoietic malignancies, including acute myeloid leukemia (AML). An optimized inhibitor (WC36) specifically and potently suppressed oncogenic activities of Syk in AML cell lines and patient-derived AML cells. Unlike ATP-competitive Syk inhibitors, WC36 was refractory to de novo and acquired drug resistance due to its ability to block not only the Syk kinase activity, but also its noncatalytic scaffolding function that is linked to drug resistance. Collectively, our study shows that targeting lipid–protein interaction is a powerful approach to developing new small molecule drugs.

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: The membrane-binding mechanisms of SH2 domains.
Fig. 2: Small molecule inhibition of the cellular activity of Syk, BLNK and PLCγ2.
Fig. 3: Inhibition of the signaling activity and proliferation of MV4-11 AML cells by Syk-cSH2 inhibitors.
Fig. 4: Suppression of HL-60 and patient-derived AML cells by Syk-cSH2 inhibitors.
Fig. 5: Signaling function of Syk in entospletinib-resistant MV4-11 cells determined by genetic modulation.
Fig. 6: Cellular target validation of WC36.

Similar content being viewed by others

Data availability

All described data are contained within the paper and its Supplementary Information files. The raw data will be shared upon request to the corresponding author. Source data are provided with this paper.

References

  1. Cho, W. & Stahelin, R. V. Membrane-protein interactions in cell signaling and membrane trafficking. Annu. Rev. Biophys. Biomol. Struct. 34, 119–151 (2005).

    Article  CAS  Google Scholar 

  2. Saliba, A. E., Vonkova, I. & Gavin, A. C. The systematic analysis of protein-lipid interactions comes of age. Nat. Rev. Mol. Cell Biol. 16, 753–761 (2015).

    Article  CAS  Google Scholar 

  3. Di Paolo, G. & De Camilli, P. Phosphoinositides in cell regulation and membrane dynamics. Nature 443, 651–657 (2006).

    Article  Google Scholar 

  4. Kihara, Y. Introduction: druggable lipid signaling pathways. Adv. Exp. Med Biol. 1274, 1–4 (2020).

    Article  CAS  Google Scholar 

  5. Nawrotek, A. et al. PH-domain-binding inhibitors of nucleotide exchange factor BRAG2 disrupt Arf GTPase signaling. Nat. Chem. Biol. 15, 358–366 (2019).

    Article  CAS  Google Scholar 

  6. Burgett, A. W. et al. Natural products reveal cancer cell dependence on oxysterol-binding proteins. Nat. Chem. Biol. 7, 639–647 (2011).

    Article  CAS  Google Scholar 

  7. Meuillet, E. J. Novel inhibitors of AKT: assessment of a different approach targeting the pleckstrin homology domain. Curr. Med. Chem. 18, 2727–2742 (2011).

    Article  CAS  Google Scholar 

  8. Lemmon, M. A. Membrane recognition by phospholipid-binding domains. Nat. Rev. Mol. Cell Biol. 9, 99–111 (2008).

    Article  CAS  Google Scholar 

  9. Muller, M. P. et al. Characterization of lipid-protein interactions and lipid-mediated modulation of membrane protein function through molecular simulation. Chem. Rev. 119, 6086–6161 (2019).

    Article  CAS  Google Scholar 

  10. Pawson, T. & Nash, P. Assembly of cell regulatory systems through protein interaction domains. Science 300, 445–452 (2003).

    Article  CAS  Google Scholar 

  11. Pawson, T. Specificity in signal transduction: from phosphotyrosine-SH2 domain interactions to complex cellular systems. Cell 116, 191–203 (2004).

    Article  CAS  Google Scholar 

  12. Lim, W. A. & Pawson, T. Phosphotyrosine signaling: evolving a new cellular communication system. Cell 142, 661–667 (2010).

    Article  CAS  Google Scholar 

  13. Liu, B. A. & Nash, P. D. Evolution of SH2 domains and phosphotyrosine signalling networks. Philos. Trans. R. Soc. Lond. B Biol. Sci. 367, 2556–2573 (2012).

    Article  CAS  Google Scholar 

  14. Park, M. J. et al. SH2 domains serve as lipid-binding modules for pTyr-signaling proteins. Mol. Cell 62, 7–20 (2016).

    Article  CAS  Google Scholar 

  15. Sheng, R. et al. Lipids regulate lck protein activity through their interactions with the lck src homology 2 domain. J. Biol. Chem. 291, 17639–17650 (2016).

    Article  CAS  Google Scholar 

  16. Kim, E. et al. Cellular phosphatase activity of C1-Ten/Tensin2 is controlled by phosphatidylinositol-3,4,5-triphosphate binding through the C1-Ten/Tensin2 SH2 domain. Cell Signal 51, 130–138 (2018).

    Article  CAS  Google Scholar 

  17. Kraskouskaya, D., Duodu, E., Arpin, C. C. & Gunning, P. T. Progress towards the development of SH2 domain inhibitors. Chem. Soc. Rev. 42, 3337–3370 (2013).

    Article  CAS  Google Scholar 

  18. Levitzki, A. Tyrosine kinase inhibitors: views of selectivity, sensitivity, and clinical performance. Annu. Rev. Pharmacol. Toxicol. 53, 161–185 (2013).

    Article  CAS  Google Scholar 

  19. Roskoski, R. Jr. Classification of small molecule protein kinase inhibitors based upon the structures of their drug-enzyme complexes. Pharmacol. Res. 103, 26–48 (2016).

    Article  CAS  Google Scholar 

  20. Roskoski, R. Jr. Properties of FDA-approved small molecule protein kinase inhibitors. Pharmacol. Res. 144, 19–50 (2019).

    Article  CAS  Google Scholar 

  21. Mocsai, A., Ruland, J. & Tybulewicz, V. L. The SYK tyrosine kinase: a crucial player in diverse biological functions. Nat. Rev. Immunol. 10, 387–402 (2010).

    Article  CAS  Google Scholar 

  22. Liu, D. & Mamorska-Dyga, A. Syk inhibitors in clinical development for hematological malignancies. J. Hematol. Oncol. 10, 145 (2017).

    Article  Google Scholar 

  23. Sharman, J. & Di Paolo, J. Targeting B-cell receptor signaling kinases in chronic lymphocytic leukemia: the promise of entospletinib. Ther. Adv. Hematol. 7, 157–170 (2016).

    Article  CAS  Google Scholar 

  24. Cremer, A. et al. Resistance mechanisms to SYK inhibition in acute myeloid leukemia. Cancer Disco. 10, 214–231 (2020).

    Article  CAS  Google Scholar 

  25. Wen, P. C. et al. Microscopic view of lipids and their diverse biological functions. Curr. Opin. Struct. Biol. 51, 177–186 (2018).

    Article  CAS  Google Scholar 

  26. Ohkubo, Y. Z., Pogorelov, T. V., Arcario, M. J., Christensen, G. A. & Tajkhorshid, E. Accelerating membrane insertion of peripheral proteins with a novel membrane mimetic model. Biophys. J. 102, 2130–2139 (2012).

    Article  CAS  Google Scholar 

  27. Soubias, O. et al. Membrane surface recognition by the ASAP1 PH domain and consequences for interactions with the small GTPase Arf1. Sci. Adv. 6, eabd1882 (2020).

    Article  CAS  Google Scholar 

  28. Narula, S. S. et al. Solution structure of the C-terminal SH2 domain of the human tyrosine kinase Syk complexed with a phosphotyrosine pentapeptide. Structure 3, 1061–1073 (1995).

    Article  CAS  Google Scholar 

  29. Yasuda, T. & Yamamoto, T. Analysis of B-cell signaling using DT40 B-cell line. Methods Mol. Biol. 271, 261–270 (2004).

    CAS  Google Scholar 

  30. Liu, S. L. et al. Quantitative lipid imaging reveals a new signaling function of phosphatidylinositol-3,4-bisphophate: isoform- and site-specific activation of Akt. Mol. Cell 71, 1092–1104 e5 (2018).

    Article  CAS  Google Scholar 

  31. DiNitto, J. P. & Lambright, D. G. Membrane and juxtamembrane targeting by PH and PTB domains. Biochim. Biophys. Acta 1761, 850–867 (2006).

    Article  CAS  Google Scholar 

  32. Chernyak, D., Chernyak, N. & Gevorgyan, V. Efficient and general synthesis of 3-aminoindolines and 3-aminoindoles via copper-catalyzed three component coupling reaction. Adv. Synth. Catal. 352, 961–966 (2010).

    Article  CAS  Google Scholar 

  33. Chernyak, N. & Gevorgyan, V. General and efficient copper-catalyzed three-component coupling reaction towards imidazoheterocycles: one-pot synthesis of alpidem and zolpidem. Angew. Chem. Int. Ed. Engl. 49, 2743–2746 (2010).

    Article  CAS  Google Scholar 

  34. Kim, H., Afsari, H. S. & Cho, W. High-throughput fluorescence assay for membrane-protein interaction. J. Lipid Res. 54, 3531–3538 (2013).

    Article  CAS  Google Scholar 

  35. Cho, W. Building signaling complexes at the membrane. Sci. STKE 2006, pe7 (2006).

    Article  Google Scholar 

  36. Wang, H. et al. ZAP-70: an essential kinase in T-cell signaling. Cold Spring Harb. Perspect. Biol. 2, a002279 (2010).

    Article  Google Scholar 

  37. Döhner, H., Weisdorf, D. J. & Bloomfield, C. D. Acute myeloid leukemia. N. Engl. J. Med. 373, 1136–1152 (2015).

    Article  Google Scholar 

  38. Yang, X. & Wang, J. Precision therapy for acute myeloid leukemia. J. Hematol. Oncol. 11, 3 (2018).

    Article  Google Scholar 

  39. Cioccio, J. & Claxton, D. Therapy of acute myeloid leukemia: therapeutic targeting of tyrosine kinases. Expert Opin. Investig. Drugs 28, 337–349 (2019).

    Article  CAS  Google Scholar 

  40. Fernandez, S. et al. Targeting tyrosine kinases in acute myeloid leukemia: why, who and how? Int. J. Mol. Sci. 20, 3429 (2019).

    Article  CAS  Google Scholar 

  41. Oellerich, T. et al. beta2 integrin-derived signals induce cell survival and proliferation of AML blasts by activating a Syk/STAT signaling axis. Blood 121, S1–S66 (2013).

    Article  Google Scholar 

  42. Bartaula-Brevik, S., Lindstad Brattas, M. K., Tvedt, T. H. A., Reikvam, H. & Bruserud, O. Splenic tyrosine kinase (SYK) inhibitors and their possible use in acute myeloid leukemia. Expert Opin. Investig. Drugs 27, 377–387 (2018).

    Article  CAS  Google Scholar 

  43. Puissant, A. et al. SYK is a critical regulator of FLT3 in acute myeloid leukemia. Cancer Cell 25, 226–242 (2014).

    Article  CAS  Google Scholar 

  44. Lange, B. et al. Growth factor requirements of childhood acute leukemia: establishment of GM-CSF-dependent cell lines. Blood 70, 192–199 (1987).

    Article  CAS  Google Scholar 

  45. Law, C. L., Chandran, K. A., Sidorenko, S. P. & Clark, E. A. Phospholipase C-gamma1 interacts with conserved phosphotyrosyl residues in the linker region of Syk and is a substrate for Syk. Mol. Cell. Biol. 16, 1305–1315 (1996).

    Article  CAS  Google Scholar 

  46. Stahelin, R. V. & Cho, W. Differential roles of ionic, aliphatic, and aromatic residues in membrane-protein interactions: a surface plasmon resonance study on phospholipases A2. Biochemistry 40, 4672–4678 (2001).

    Article  CAS  Google Scholar 

  47. Zhang, J. H., Chung, T. D. & Oldenburg, K. R. A simple statistical parameter for use in evaluation and validation of high throughput screening assays. J. Biomol. Screen 4, 67–73 (1999).

    Article  CAS  Google Scholar 

  48. Khan, I. et al. FOXM1 contributes to treatment failure in acute myeloid leukemia. JCI Insight 3, e121583 (2018).

    Article  Google Scholar 

  49. Wells, D. B., Abramkina, V. & Aksimentiev, A. Exploring transmembrane transport through alpha-hemolysin with grid-steered molecular dynamics. J. Chem. Phys. 127, 125101 (2007).

    Article  Google Scholar 

  50. Phillips, J. C. et al. Scalable molecular dynamics with NAMD. J. Comput. Chem. 26, 1781–1802 (2005).

    Article  CAS  Google Scholar 

  51. Roos, K. et al. OPLS3e: extending force field coverage for drug-like small molecules. J. Chem. Theory Comput. 15, 1863–1874 (2019).

    Article  CAS  Google Scholar 

  52. Sastry, G. M., Adzhigirey, M., Day, T., Annabhimoju, R. & Sherman, W. Protein and ligand preparation: parameters, protocols, and influence on virtual screening enrichments. J. Comput Aided Mol. Des. 27, 221–234 (2013).

    Article  Google Scholar 

  53. Halgren, T. A. Identifying and characterizing binding sites and assessing druggability. J. Chem. Inf. Model. 49, 377–389 (2009).

    Article  CAS  Google Scholar 

  54. Friesner, R. A. et al. Extra precision glide: docking and scoring incorporating a model of hydrophobic enclosure for protein-ligand complexes. J. Med. Chem. 49, 6177–6196 (2006).

    Article  CAS  Google Scholar 

  55. Zhang, J. et al. Identifying mutation hotspots reveals pathogenetic mechanisms of KCNQ2 epileptic encephalopathy. Sci. Rep. 10, 4756 (2020).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This work was supported by grants from the National Institutes of Health (grant nos. R35GM122530 to W.C., GM 120281 to V.G., P41-GM104601 and R01-GM123455 to E.T. and R01-NS114413 to S.C.). I.K. acknowledges University of Illinois Cancer Center’s 2020 Pilot Program (grant no. 2020-PP-01). E.T. thanks Blue Waters at National Center for Supercomputing Applications for providing computing resources and Extreme Science and Engineering Discovery Environment (grant no. MCA06N060). E.T. is also grateful to the School of Chemical Sciences Scientific Software Program for access to the Schrödinger Suite.

Author information

Authors and Affiliations

Authors

Contributions

A.S. and I.S. carried out cell studies on Syk and Syk inhibitors. A.S. performed synthesis, screening, optimization and characterization of inhibitors. S.P., M.L. and E.T. performed computational analysis. M.-J.P. and P.B. performed SPR studies on Syk-SH2 and BLNK-SH2, respectively. Y.H. contributed to initial Syk inhibitor screening. N.M. and I.K. collected and performed experiments with the clinical samples. M.P. and S.C. performed MS analysis. V.G. constructed a small molecule library. Y.-M.K. participated in BCR signaling analysis. W.C. conceived and supervised the work and wrote the paper.

Corresponding author

Correspondence to Wonhwa Cho.

Ethics declarations

Competing interests

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest. S.P. is currently an employee of Loxo Oncology @ Lilly and is a shareholder of stock in Eli Lilly and Co.

Peer review

Peer review information

Nature Chemical Biology thanks Thomas Smithgall and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

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

Extended data

Extended Data Fig. 1 PIP3-dependent membrane binding of Syk-cSH2.

a. A representative simulation box for the HMMM membrane binding simulations of Syk-cSH2, which is initially placed at least 10 Å away from the membrane surface. 1,2-dihexanoyl derivatives of phosphatidylcholine (PC) and phosphatidylserine (PS) are shown in gray and purple lines, respectively, and PIP3 (6 in each leaflet) in space-filling representation (carbon, oxygen, and phosphorous atoms in yellow, red and tan, respectively). The hydrophobic core of the membrane is filled with an organic solvent (DCLE), shown in transparent surface representation. Bulk water molecules and ions are not shown for clarity. b. Ensemble-averaged residue profile of the Syk-cSH2, calculated during last the 50 ns of HMMM simulations. The gray bars denote the standard deviation in the residue distances. The pY-interacting residue (R195) is highlighted in black star. The representative PIP3-interacting conformations captured at the end of membrane binding simulations are shown at the bottom panel. c. Probability histogram of the number of PIP3 within 5 Å of Syk-cSH2, calculated over last 50 ns of all the HMMM membrane binding simulations, highlights that the protein can simultaneously interact with multiple PIP3 interacting which might help in its effective recruitment. d. Histograms of the protein residues in contact (within 3.5 Å) with the 3’, 4’ and 5’ phosphate moieties (P3 (red), P4 (green), and P5 (blue)) of the PIP3 headgroup are calculated over last 50 ns of all the HMMM membrane binding simulations. Notice that K220 and K222 make contact primarily with P3 whereas H163 and K165 interact with both P3 and P4. Many other residues primarily interact with P5. e. All the residues that form contacts with PIP3 for more than 10% of the time are highlighted and those that make the closest contact with PIP3 are labeled. All the basic residues are colored in blue and polar residues in green. f. Enlarged view of interaction between K172 of the Syk-cSH2 domain and the P4 and P5 moieties of PIP3. g. A schematic representation of predicted hydrogen bonds in this binding mode. H-bonds are shown as arrows. Notice that K172 interacts with both P4 and P5, but not with P3. It is thus expected to be involved in non-specific interaction with non-3’ P-containing PtdInsPs, such as PI(4,5)P2. h. Determination of Kd for binding of Syk-cSH2 WT (blue) and mutants (K220A/K222A (red), K172A (green), and R195A (orange) to 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC)/ 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoserine (POPS)/PIP3 (77:20:3) large unilamellar vesicles (LUVs). The protein concentration was varied from 0 to 500 nM. Each data is average ± SD from 3 independent measurements.

Extended Data Fig. 2 Role of Syk-cSH2-PIP3 binding in plasma membrane translocation of Syk.

a. Subcellular localization of EGFP-Syk-WT and a PIP3 sensor, mCherry-eMyoX-tPH co-expressed in Syk-null DT40 cells after IgM stimulation. b. Subcellular localization of EGFP-Syk-WT and mCherry-eMyoX-tPH expressed in Syk-null DT40 cells before IgM stimulation. c. The effect of depleting cellular PIP3 on IgM-stimulated membrane translocation of EGFP-Syk-WT and mCherry-eMyoX-tPH was measured by treating cells with a Class I PI3K inhibitor, GDC-0941 (10 nM, overnight). d. Subcellular localization of EGFP-Syk-K220A/K222A and mCherry-eMyoX-tPH expressed in Syk-null DT40 cells after IgM stimulation. IgM stimulation was performed by dropping cells on the IgM-coated surface (10 ng/ml) for 10 min. r values indicate Pearson’s correlation coefficients (average ± SD from 3 independent measurements). r was not determined (ND) for Extended Data Fig. 2c because localization of neither EGFP-Syk-WT nor mCherry-eMyoX-tPH was observed in the plasma membrane. Scale bars indicate 2.5 μm. e. IgM-stimulated (10 ng/ml for 10 min) Ca2+ release in Syk-null DT40 cells (black line) stably transfected with Syk-WT (blue line) and Syk-K220A/K222A (red line). Ca2+ levels were ratiometrically (F340/F380: the ratio of fluorescence emission intensity at 380 nm to that at 340 nm) estimated using Fura-2. Each data is a representative curve of 3 independent measurements. Inset: expression levels of transfected Syk-WT and K220A/K222A were comparable as seen in the similar density of immunoblot bands. Glyceraldehyde 3-phosphate dehydrogenase (GAPDH) was used as a gel loading control. The image is representative of 3 independent measurements.

Source data

Extended Data Fig. 3 PIP3-dependent membrane binding of BLNK-SH2.

a. A representative simulation box for the HMMM membrane binding simulations of BLNK-SH2, which is initially placed at least 10 Å away from the membrane surface. See Extended Data Fig. 1 for details. b. A predicted mode of binding of BLNK-SH2 to a PIP3 molecule in the membrane. Left panel: a PIP3 molecule (C and O, N atoms are shown in yellow, red and blue, respectively) and BLNK-SH2 residues (C and O, N atoms are shown in white, red and blue, respectively) interacting with it are shown in stick presentation and labeled. Gray circles indicate PC and PS molecules. This representative structure is one of multiple low energy conformations obtained from the simulation and docking and not all of residues predicted for PIP3 binding are shown. Right panel: the structure is rotated 90o for the top of the protein surface to face the membrane. The protein is shown in surface representation to better illustrate the shape of the primary lipid binging site. c. Superimposition of PIP3 and VG594 docked into the primary lipid binding groove of BLNK-SH2 in the membrane environment. In its lowest energy binding mode, VG594 binds deeper to the groove than the PIP3 headgroup, leading to tight binding and effective inhibition of PIP3 binding. d. Determination of Kd for binding of BLNK-SH2 WT (circles) and K411A/K412A/K413A (triangles) to POPC/POPS/PIP3 (77:20:3) LUVs. The protein concentration was varied from 0 to 500 nM. Each data is average ± SD from 3 independent measurements. e. Selectivity of BLNK-SH2 WT (red) and K411A/K412A/K413A (blue) for POPC/POPS/ PIP3 (77:20:3) (solid lines) over POPC/POPS/PI(4,5)P2 (PIP2) (77:20:3) (broken lines) vesicles determined by SPR analysis. The protein concentration was 100 nM.

Extended Data Fig. 4 PIP3-dependent membrane binding of PLCγ2-cSH2.

a. A representative simulation box for the HMMM membrane binding simulations of PLCγ2-cSH2, which is initially placed at least 10 Å away from the membrane surface. See Extended Data Fig. 1 for details. b. A predicted mode of binding of PLCγ2-cSH2 to a PIP3 molecule in the membrane. Left panel: a PLCγ2-cSH2 residues (C and O, N atoms are shown in white, red and blue, respectively) interacting with a PIP3 molecule (C and O, N atoms are shown in yellow, red and blue, respectively) are shown in stick presentation and labeled. Gray circles indicate PC and PS molecules. This representative structure is one of multiple low energy conformations obtained from the simulation and docking and not all of residues predicted for PIP3 binding are shown. Right panel: the structure is rotated 90° for the top of the protein surface to face the membrane. The protein is shown in surface representation to better illustrate the shape of the primary lipid binging site. c. Superimposition of PIP3 and VG370 docked into the primary lipid binding groove of PLCγ2-cSH2 in the membrane environment. In its lowest energy binding mode, VG370 binds deeper to the groove than the PIP3 headgroup, leading to tight binding and effective inhibition of PIP3 binding. d. Determination of Kd for binding of PLCγ2-cSH2 WT (circles) and K727A/K728A (triangles) to POPC/POPS/PIP3 (77:20:3) LUVs. The protein concentration was varied from 0 to 500 nM. Each data is average ± SD from 3 independent measurements. (e) Selectivity of PLCγ2-cSH2 WT (red) and K727A/K728A (blue) for POPC/POPS/PIP3 (77:20:3) (solid lines) over POPC/POPS/PI(4,5)P2 (PIP2) (77:20:3) (broken lines) vesicles determined by SPR analysis. The protein concentration was 100 nM.

Extended Data Fig. 5 Efficacy and specificity of Syk inhibitors.

a. Cellular efficacy of WC35, WC36, and WC38 was determined by dose-dependent inhibition of Syk phosphorylation in Raji B cells. b. Imax and IC50 values (see Supplementary Table 3) were determined by nonlinear least-squares analysis of data using the equation: I = Imax / (1 + IC50 / [inhibitor]) where I indicates %inhibition. %inhibition was calculated by 100 x (relative decrease in pSyk/(total Syk) at a given inhibitor concentration ([inhibitor]) compared to that without an inhibitor. Each data represents average ± SD from 3 independent measurements. c. Specificity of WC35, WC36, WC38, and entospletinib (ENTO) was determined by measuring the inhibition of ZAP70 phosphorylation by these molecules (5 μM overnight) in Jurkat T cells. Jurkat T cells were stimulated by OKT3. Each gel image is a representative of 3 independent experiments.

Source data

Extended Data Fig. 6 Proposed mechanisms of Syk signaling, inhibition, and resistance in AML cells.

a. In the resting state, Syk is autoinhibited by intramolecular tethering by interdomain A (ID-A; dark red) and B (ID-B) which occludes the kinase (Kin) domain. b. IgG2-activated FcγRI causes the associated ITAM motif to be doubly phosphorylated (pY2-ITAM) by a Src family kinase, creating a docking site for Syk. c. Recruitment of Syk to the activated receptor requires coincident Syk-cSH2-PIP3 binding and Syk-n/cSH2-pY2-ITAM binding, which leads to conformational changes and activation of Syk. d. WC36 specifically blocks this process by inhibiting Syk-cSH2-PIP3 binding while entospletinib cannot. e. Once activated by the receptor, Syk typically undergoes autophosphorylation (pY: green) of ID-A and ID-B for sustained activation and functions as both a kinase and a scaffolding protein. As a kinase, Syk phosphorylates a constitutively active FLT3 mutant (for example, FLT3-ITD), which is essential for AML cell proliferation. Activated Syk also serves as a scaffolding protein for FLT3, RAS-RAF-MEK-ERK and STA3/5 either directly or via an adaptor protein (for example, X = KSR). f. Entospletinib can block the kinase activity of Syk but cannot inhibit the scaffolding function, allowing the AML cells to develop acquired resistance mechanism that bypasses the Syk kinase activity. WC36 does not allow any type of resistance because it blocks all functions of Syk at the source.

Extended Data Fig. 7 Target validation of WC36.

a. Design of mutants that may potentially inhibit WC36 binding without affecting PIP3 binding. In the model structures of Syk-cSH2-PIP3 and Syk-cSH2-WC36 complexes (superimposed here as in Fig. 2d), the side chain of E164 forms a hydrogen bond (the red dotted line) with the N-H group of the imidazole moiety of WC36 but is remote from the PIP3 head group. We thus reasoned that the mutation of E164 might selectively suppress WC36 binding without impacting PIP3 binding. The blue dotted line indicates the location of the membrane surface. b. Membrane binding activity of EGFP-Syk-cSH2-WT (blue), -E164D (red), -E164A (purple), -E164Q (green), and -E164K (orange) was measured by the fluorescence quenching assay using POPC/POPS/PIP3/dabsyl-PE (67:20:3:10 in mol%) LUVs. The quenching of the EGF fluorescence by dabsyl-PE (F) of was determined as a function of the total lipid concentration in the vesicles. E164D showed essentially the same vesicle affinity as WT. The protein concentration was kept at 25 nM. c. Inhibition of membrane binding activity of EGFP-Syk-cSH2-WT (blue), -E164D (red), -E164A (purple), -E164Q (green), and -E164K (orange) by WC36. 25 nM of Syk-cSH2 proteins and 50 μM POPC/POPS/PIP3/dabsyl-PE (67:20:3:10 in mol%) LUVs were employed. % inhibition ((F / Fmax) x 100)) was plotted as the function of the WC36 concentration where F and Fmax indicate fluorescence quenching in the presence of and absence of WC36, respectively. Neither E164D nor W164Q was inhibited by WC36 whereas WT was. For Expended Data Fig. 7b, d, each data represents average ± SD from 3 independent measurements. d. The structure of WC36B.

Supplementary information

Supplementary Information

Supplementary Figs. 1–7, Tables 1–5, Notes 1 and 2 and Data files 1 and 2.

Reporting Summary

Supplementary Data 1

Proteins captured by WC36B and biotin were subject to identification by data-dependent mass spectrometry analysis. Raw data was searched against the Swiss Protein Homosapien database using the Proteome Discoverer software using the Sequest HT algorithm. Shown herein are proteins identified in the capture by WC36B and biotin. In this analysis, Syk as only detected in the WC36B fluorophore capture. Peptide sequences for the 28 unique peptides (47% sequence coverage) for Syk are also included. In our approach, the more intense a peptide, the more it is selected for fragmentation (peptide spectral matches, PSM). Although peptides may have different ionization efficiencies, PSM values can be in general related to the abundance of proteins in the capture. In the WC36B-captured proteins, Syk has 219 spectra whereas the next highest one is 79, 71, 61, 15, 11 and many at only 2 PSMs. Many of these proteins were also identified in the biotin capture, whereas not a single peptide was identified for Syk in the biotin capture. Collectively, these results show Syk is the major protein species captured by WC36B.

Source data

Source Data Fig. 2

Unprocessed western blots.

Source Data Fig. 3

Unprocessed western blots.

Source Data Fig. 4

Unprocessed western blots.

Source Data Fig. 5

Unprocessed western blots.

Source Data Fig. 6

Unprocessed western blots.

Source Data Extended Data Fig. 2

Unprocessed western blots.

Source Data Extended Data Fig. 5

Unprocessed western blots.

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Singaram, I., Sharma, A., Pant, S. et al. Targeting lipid–protein interaction to treat Syk-mediated acute myeloid leukemia. Nat Chem Biol 19, 239–250 (2023). https://doi.org/10.1038/s41589-022-01150-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41589-022-01150-z

This article is cited by

Search

Quick links

Nature Briefing: Cancer

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

Get what matters in cancer research, free to your inbox weekly. Sign up for Nature Briefing: Cancer