Methylation of dual-specificity phosphatase 4 controls cell differentiation

Hairui Su, Ming Jiang, Chamara Senevirathne, Srinivas Aluri, Tuo Zhang, Han Guo, Juliana Xavier-Ferrucio, Shuiling Jin, Ngoc-Tung Tran, Szu-Mam Liu, Chiao-Wang Sun, Yongxia Zhu, Qing Zhao, Yuling Chen, LouAnn Cable, Yudao Shen, Jing Liu, Cheng-Kui Qu, Xiaosi Han, Christopher A Klug, Ravi Bhatia, Yabing Chen, Stephen D Nimer, Y George Zheng, Camelia Iancu-Rubin, Jian Jin, Haiteng Deng, Diane S Krause, Jenny Xiang, Amit Verma, Minkui Luo, Xinyang Zhao, Hairui Su, Ming Jiang, Chamara Senevirathne, Srinivas Aluri, Tuo Zhang, Han Guo, Juliana Xavier-Ferrucio, Shuiling Jin, Ngoc-Tung Tran, Szu-Mam Liu, Chiao-Wang Sun, Yongxia Zhu, Qing Zhao, Yuling Chen, LouAnn Cable, Yudao Shen, Jing Liu, Cheng-Kui Qu, Xiaosi Han, Christopher A Klug, Ravi Bhatia, Yabing Chen, Stephen D Nimer, Y George Zheng, Camelia Iancu-Rubin, Jian Jin, Haiteng Deng, Diane S Krause, Jenny Xiang, Amit Verma, Minkui Luo, Xinyang Zhao

Abstract

Mitogen-activated protein kinases (MAPKs) are inactivated by dual-specificity phosphatases (DUSPs), the activities of which are tightly regulated during cell differentiation. Using knockdown screening and single-cell transcriptional analysis, we demonstrate that DUSP4 is the phosphatase that specifically inactivates p38 kinase to promote megakaryocyte (Mk) differentiation. Mechanistically, PRMT1-mediated methylation of DUSP4 triggers its ubiquitinylation by an E3 ligase HUWE1. Interestingly, the mechanistic axis of the DUSP4 degradation and p38 activation is also associated with a transcriptional signature of immune activation in Mk cells. In the context of thrombocytopenia observed in myelodysplastic syndrome (MDS), we demonstrate that high levels of p38 MAPK and PRMT1 are associated with low platelet counts and adverse prognosis, while pharmacological inhibition of p38 MAPK or PRMT1 stimulates megakaryopoiesis. These findings provide mechanistic insights into the role of the PRMT1-DUSP4-p38 axis on Mk differentiation and present a strategy for treatment of thrombocytopenia associated with MDS.

Trial registration: ClinicalTrials.gov NCT01496495.

Keywords: DUSP4; HUWE1; MDS; PRMT1; leukemia; megakaryocyte; myelodysplasia syndrome; p38; platlet; trombocytopenia.

Conflict of interest statement

Declaration of interests M.L. has served on the Scientific Advisory Board for Epi One. A.V. has received research funding from GlaxoSmithKline, Incyte, MedPacto, Novartis, Curis, and Eli Lilly and Company; has received compensation as a scientific advisor to Novartis, Stelexis Therapeutics, Acceleron Pharma, and Celgene; and has equity ownership in Stelexis Therapeutics. The remaining authors declare no competing interests. Array BioPharma provided the p38 inhibitor pexmetinib (ARRY614) and participated in its phase I study (ClinicalTrials.gov: NCT01496495).

Copyright © 2021 The Authors. Published by Elsevier Inc. All rights reserved.

Figures

Figure 1.. Identification of DUSP4 for optimal…
Figure 1.. Identification of DUSP4 for optimal Mk differentiation
(A) Schematic of shRNA-based screening assay to identify essential DUSPs for Mk-induced Mk differentiation. Human CD34+ cells infected with lentiviruses expressing shRNAs against DUSPs were cultured for Mk differentiation. (B) Heatmap of the percentages of CD41a+CD42b+ cells upon DUSP knockdown on day 8. Fold changes were normalized to the percentage of double-positive cells with the group treated with control shRNA. (C) Representative flow chart of FACS analysis of Mk differentiation using BM cells (top panel) and CB cells (bottom panel) cultured in TPO-containing medium. (D) Summary of FACS analysis. Statistics are based on the data of three independent experiments (n = 3) with the bone marrow (BM) or cord blood (CB) cells from two donors. Data are shown as mean ± SD. Two-tailed paired t test, *p ≤ 0.05, **p ≤ 0.01. (E) Schematic description of Mk differentiation of human peripheral blood-derived CD34+ cells with DUSP4 knockdown. (F and G) Representative and complete FACS analysis of CD41a and CD42b markers for Mk differentiation with peripheral blood CD34+ cells upon DUSP4 knockdown on day 7. Representative plots (F) and statistics (G) are shown (n = 3, independent experiments). Data are shown as mean ± SD. Two-tailed paired t test, *p ≤ 0.05, **p ≤ 0.01.
Figure 2.. DUSP4-regulated differentiation choices between Mk…
Figure 2.. DUSP4-regulated differentiation choices between Mk cells and Er cells
(A) Schematic of differentiation experiments using CB CD34+ cells. (B) FACS analysis of CD41a and CD71 on the cultured cells. Representative plots (top panel) and normalized statistics of three samples (bottom panel) are shown (n = 3, independent experiments). Fold changes were normalized with scramble controls. (C) Schematic description of colony-forming unit (CFU) assays using MEP. (D) Percentages of CFU-Mk, BFU-E, and CFU-MkE in each sorted population (n = 2, independent experiments).
Figure 3.. Crosstalk between DUSP4 and PRMT1…
Figure 3.. Crosstalk between DUSP4 and PRMT1 for MAPK signaling in Mk differentiation
(A) DUSP4 mRNA level in PMA-treated MEG-01 cells. Cells were harvested at indicated time intervals, and the extracted mRNAs were quantified by real-time PCR. Representative statistics are shown as mean ± SD. Two-tailed unpaired t test, **p ≤ 0.01, ***p ≤ 0.001 (n = 3, independent experiments). (B) MAPK-related proteins and PRMT1 in MEG-01 cells after PMA stimulation. Protein extracts were collected at indicated time intervals for western blotting. (C) Regulation of MAPK signaling upon DUSP4 overexpression. MEG-01 cells were treated overnight with doxycycline to induce DUSP4 ectopically expressed from lentivirus. Cell extracts were collected for western blotting (n = 3, representative western blots). (D) PRMT1 mRNA level in MEG-01 cells during the course of PMA-stimulated Mk differentiation. Representative statistics were shown as mean ± SD, two-tailed unpaired t test, *p ≤ 0.05 (n = 3, independent experiments). (E and F) PRMT1-dependent regulation of DUSP4 protein. NB4 cells that conditionally express shRNA against PRMT1 were treated with doxycycline to induce PRMT1 knockdown (E). DUSP4- and PRMT1-encoding plasmids were transfected into HEK293T cells for their overexpression in the presence or absence of MG132 treatment (F). (G) Antagonistic roles of PRMT1 and DUSP4 on Mk differentiation of human CD34+ cells. Human CD34+ cells were infected with PRMT1 lentivirus (puromycin-R) and DUSP4 lentivirus (GFP), followed by puromycin selection and Mk differentiation. Representative plots and statistics are shown (n = 3, independent experiments). Data are shown as mean ± SD. Two-tailed paired t test, *p ≤ 0.05, **p ≤ 0.01.
Figure 4.. Negative correlation between PRMT1 and…
Figure 4.. Negative correlation between PRMT1 and Mk differentiation revealed by single-cell RNA-seq (scRNA-seq) analysis
(A) Experimental design for scRNA-seq analysis. Sample 1: Native CD34+ cells were cells isolated directly from BM. Sample 2: CD34+ cells were cells cultured in TPO and SCF for 8 days before sorting with CD41a and CD42b. Sample 3: The non-CD41a+CD42b+ cells were sorted from sample 2. Sample 4: The CD41a+CD42b+ cells were sorted from sample 2. (B and C) SPRING plots of single-cell transcriptomes. Individual cells are presented according to their origins (B) or transcriptome-associated cell types (C). Ba, basophilic or mast cell; D, dendritic; Er, erythroid; GN, granulocytic neutrophil; Ly, lymphocytic; M, monocytic; Mk, megakaryocytic; MPP, multipotential progenitor. (D) Pathway analysis of the top 400 genes with the strongest negative Spearman correlation to PRMT1 expression level in the TPO/SCF-stimulated CD41a+CD42b+ population. Red bars highlight Mk-relevant biological pathways. (E) Spearman correlation coefficients between PRMT1 and any gene revealed by scRNA-seq in TPO/SCF-stimulated CD41a+CD42b+ cells. Representative genes with significant correlation and functional relevance are annotated. (F) Gene set enrichment analysis (GSEA) of PRMT1-correlated genes in the TPO/SCF-stimulated CD41a+CD42b+ population. GSEA inputs are Spearman correlation coefficients of PRMT1 versus any gene with single-cell resolution. (G) Normalized expression of representative transcripts co-plotted against that of PRMT1 transcript in TPO/SCF-stimulated CD41a+CD42b+ cells with single-cell resolution.
Figure 5.. R351 methylation of DUSP4 by…
Figure 5.. R351 methylation of DUSP4 by PRMT1
(A) Schematic of the next-generation live-cell BPPM technology to uncover substrates of PRMT1. (B) Immunoblotting readouts of H4 and DUSP4 as PRMT1 targets enriched via the BPPM technology using HEK293T cells (n = 3, representative western blots). (C) Schematic description of the next-generation live-cell BPPM technology to reveal methylation sites of PRMT1. Candidates of methylated arginine (Arg) residues on a protein substrate are mutated to lysine (Lys). The resulting arginine-to-lysine mutation is expected to diminish or abolish the BPPM-associated labeling. (D) DUSP4 sequence with functional domains and RXR motifs highlighted. (E and F) Revealing PRMT1 methylation sites on DUSP4 with the next-generation live-cell BPPM technology. The DUSP4 variants contain dual arginine-to-lysine mutations at RXR motifs (E) and the point mutations at R351 and R353 (F), respectively (n = 3, representative western blots). (G) Validating PRMT1-invovled R351 methylation on DUSP4 with anti-methylarginine antibodies in HEK293T cells (n = 3, representative western blots). Two anti-methyl-arginine antibodies were used to determine the presence of arginine methylation in DUSP4.
Figure 6.. Polyubiquitylation and instability of DUSP4…
Figure 6.. Polyubiquitylation and instability of DUSP4 promoted by PRMT1-involved R351 methylation
For a Figure360 author presentation of this figure, see https://doi.org/10.1016/j.celrep.2021.109421. (A) Polyubiquitylation of DUSP4, but not its R351K variant, is stimulated by PRMT1. 293T cells were transfected with DUSP4 (wild-type and R351K mutant), 6× His-tagged ubiquitin, and PRMT1 (V1 and V2 isoforms). Cells were treated with MG132 for 6 h prior to harvest. Fractions of pull-downs (upper panel) and inputs (bottom panel) were applied for western blotting (n = 3, representative western blots). (B) PRMT1-dependent stability levels of wild-type and R351K DUSP4 were determined in 293T cells by western blotting (n = 3, representative western blots). (C) Half-life time of wild-type and R351K DUSP4. Normalized protein stability curves are plotted in the right panel (n = 3, representative western blots). (D) Mechanistic description of DUSP4 stability modulated by PRMT1-dependent R351 methylation. R351 methylation of DUSP4 by PRMT1 triggers its polyubiquitylation and thus its degradation; R351K mutation abolishes the methylation and thus suppresses polyubiquitylation and degradation (n = 3, representative western blots). (E) Mk differentiation in the presence of wild-type and R351K DUSP4. Human CD34+ cells infected with lentiviruses expressing DUSP4s (wild-type or R351K mutant) were induced for Mk differentiation. Percentage of CD41a+CD42b+ cells was determined by FACS after 7 days. Representative plots and statistics are shown (n = 3, independent experiments). Data are shown as mean ± SD. Two-tailed paired t test, *p ≤ 0.05. (F) Co-immunoprecipitation of HUWE1 and DUSP4. Flag-tagged DUSP4 was used to immunoprecipitate myc-tagged HUWE1 in co-transfected 293T cells (n = 3, representative western blots after immunoprecipitation). (G) The protein ubiquitylation of DUSP4 is measured in 293T cells transfected with plasmids as shown on top of the gels (n = 3, representative western blots). (H) Mutant and wild-type DUSP4 were expressed together with HUWE1 shRNA in 293T cells for western blotting with respective antibodies (n = 3, representative western blots). (I) Protein ubiquitylation assays with wild-type and mutant DUSP4. DUSP4 wild-type protein and mutant protein were expressed in 293T cells transfected with or without the plasmid combination of HUWE1 and PRMT1 as indicated on the top of the gel for affinity purification with Ni-NTA beads (n = 3, representative western blots).
Figure 7.. Clinical implication and pharmacological targeting…
Figure 7.. Clinical implication and pharmacological targeting of the p38-DUSP4-PRMT1 axis in MDS
(A and B) Gene expression of PRMT1 and p38α in MDS patients and healthy donors with array-based analysis. The CD34+ HSPCs of MDS patients (N = 183) and age-matched healthy controls (N = 17) were analyzed for PRMT1 expression (A) and p38α expression (B). Data are shown as mean ± SD. Two-tailed unpaired t test, ****p ≤ 0.0001. (C) MDS cohorts as classified by low and high expression of p38α MAPK on the basis of median expression levels. The subjects with high p38α expression showed significantly lower platelet counts. Data are shown as mean ± SD. Two-tailed unpaired t test, *p ≤ 0.05. (D) Survival curves of MDS patients classified by low and high expression of p38α MAPK. The MDS patients with higher p38α expression in HSPCs showed significantly worse overall survival. *p ≤ 0.05. (E and F) Immunohistochemistry (IHC) analysis for phosphorylation-activated p38α MAPK of age-matched healthy controls and MDS BM samples from a clinical trial with the p38α inhibitor pexmetinib (ARRY614, labeled as p38i). MDS BMs showed significantly higher phospho-p38α staining in megakaryocytes (E). Representative stains are shown (F). Data are shown as mean ± SD. Two-tailed unpaired t test, **p ≤ 0.01. (G–I) Effects of the p38α inhibitor pexmetinib (labeled as p38i) on normal CD34+ cells. Normal CD34+ cells were grown in liquid culture conditions in the presence and absence of pexmetinib and analyzed for CD41 and CD42 expression by representative charts shown (G) and averaged FACS data (H) (n = 2, independent experiments). (I) Normal CD34+ cells were also grown for MegaCult assay for production of megakaryocyte colonies (n = 2, independent experiments). Data are shown as mean ± SD. Two-tailed paired t test, **p ≤ 0.05, **p ≤ 0.01. (J and K) Analysis of BM mononuclear cells (MNCs) of MDS patients with MegaCult assay for production of megakaryocyte colonies in the presence or absence of the p38α inhibitor pexmetinib (or p38i). Six MNC samples were examined (J) with the representative images shown (K). (L) Analysis of MNCs of MDS patients (n = 6) with MegaCult assay for production of megakaryocyte colonies in the presence or absence of a PRMT1 inhibitor MS023. Data are shown as mean ± SD. Two-tailed paired t test, **p ≤ 0.01. (M) Mk polyploidy and platelet count analysis in C57BL6/J mice treated with MS023. BM CD41+ cells were analyzed by FACS for polyploidy. The number of platelets in peripheral blood and MPV (mean platelet volume) were analyzed by a Hemavet machine. (N) Mechanistic description of Mk differentiation via the PRMT1-DUSP4-p38 axis. Mk progenitors undergo abnormal differentiation in MDS by upregulation of PRMT1, which leads to p38 kinase activation. The relative levels of phospho-p38 are regulated by DUSP4. DUSP4 R351 is subject to PRMT1-medidated methylation, which leads to polyubiquitylation by HUWE1 and then degradation. Collectively, the PRMT1-DUSP4-p38 axis determines generation of Mk progenitor cells and the maturation of Mk cells.

References

    1. Al-Mutairi MS, Cadalbert LC, McGachy HA, Shweash M, Schroeder J, Kurnik M, Sloss CM, Bryant CE, Alexander J, and Plevin R (2010). MAP kinase phosphatase-2 plays a critical role in response to infection by Leishmania mexicana. PLoS Pathog. 6, e1001192.
    1. Auger-Messier M, Accornero F, Goonasekera SA, Bueno OF, Lorenz JN, van Berlo JH, Willette RN, and Molkentin JD (2013). Unrestrained p38 MAPK activation in Dusp1/4 double-null mice induces cardiomyopathy. Circ. Res. 112, 48–56.
    1. Bachegowda L, Morrone K, Winski SL, Mantzaris I, Bartenstein M, Ramachandra N, Giricz O, Sukrithan V, Nwankwo G, Shahnaz S, et al. (2016). Pexmetinib: A novel dual inhibitor of Tie2 and p38 MAPK with efficacy in preclinical models of myelodysplastic syndromes and acute myeloid leuke mia. Cancer Res. 76, 4841–4849.
    1. Baldwin RM, Morettin A, Paris G, Goulet I, and Côté J (2012). Alternatively spliced protein arginine methyltransferase 1 isoform PRMT1v2 promotes the survival and invasiveness of breast cancer cells. Cell Cycle 11, 4597–4612.
    1. Balko JM, Cook RS, Vaught DB, Kuba MG, Miller TW, Bhola NE, Sanders ME, Granja-Ingram NM, Smith JJ, Meszoely IM, et al. (2012). Profiling of residual breast cancers after neoadjuvant chemotherapy identifies DUSP4 deficiency as a mechanism of drug resistance. Nat. Med. 18, 1052–1059.
    1. Baumgartner C, Toifl S, Farlik M, Halbritter F, Scheicher R, Fischer I, Sexl V, Bock C, and Baccarini M (2018). An ERK-dependent feedback mechanism prevents hematopoietic stem cell exhaustion. Cell Stem Cell 22, 879–892.e6.
    1. Blum G, Bothwell I, Islam K, and Luo M (2013). Profiling protein methylation with cofactor analog containing terminal alkyne functionality. Curr. Protoc. Chem. Biol. 5, 67–88.
    1. Blum G, Bothwell IR, Islam K, and Luo M (2013). Profiling protein methylation with cofactor analog containing terminal alkyne functionality. Curr. Protoc. Chem. Biol. 5, 67–88.
    1. Boisvert FM, Côté J, Boulanger MC, and Richard S (2003). A proteomic analysis of arginine-methylated protein complexes. Mol. Cell. Proteomics 2, 1319–1330.
    1. Boisvert FM, Hendzel MJ, Masson JY, and Richard S (2005). Methylation of MRE11 regulates its nuclear compartmentalization. Cell Cycle 4, 981–989.
    1. Caunt CJ, and Keyse SM (2013). Dual-specificity MAP kinase phosphatases (MKPs): Shaping the outcome of MAP kinase signalling. FEBS J. 280, 489–504.
    1. Chen P, Hutter D, Yang X, Gorospe M, Davis RJ, and Liu Y (2001). Discordance between the binding affinity of mitogen-activated protein kinase subfamily members for MAP kinase phosphatase-2 and their ability to activate the phosphatase catalytically. J. Biol. Chem. 276, 29440–29449.
    1. Cheng H, Zheng Z, and Cheng T (2020). New paradigms on hematopoietic stem cell differentiation. Protein Cell 11, 34–44.
    1. Chi H, and Flavell RA (2008). Acetylation of MKP-1 and the control of inflammation. Sci. Signal. 1, pe44.
    1. D’Atri LP, Etulain J, Rivadeneyra L, Lapponi MJ, Centurion M, Cheng K, Yin H, and Schattner M (2015). Expression and functionality of Toll-like receptor 3 in the megakaryocytic lineage. J. Thromb. Haemost. 13, 839–850.
    1. Debili N, Coulombel L, Croisille L, Katz A, Guichard J, Breton-Gorius J, and Vainchenker W (1996). Characterization of a bipotent erythro-megakaryocytic progenitor in human bone marrow. Blood 88, 1284–1296.
    1. Desterke C, Bilhou-Nabéra C, Guerton B, Martinaud C, Tonetti C, Clay D, Guglielmelli P, Vannucchi A, Bordessoule D, Hasselbalch H, et al.; French Intergroup of Myeloproliferative Disorders; French INSERM; European EUMNET Networks on Myelofibrosis (2011). FLT3-mediated p38-MAPK activation participates in the control of megakaryopoiesis in primary myelofibrosis. Cancer Res. 71, 2901–2915.
    1. Dobin A, Davis CA, Schlesinger F, Drenkow J, Zaleski C, Jha S, Batut P, Chaisson M, and Gingeras TR (2013). STAR: Ultrafast universal RNA-seq aligner. Bioinformatics 29, 15–21.
    1. Eram MS, Shen Y, Szewczyk M, Wu H, Senisterra G, Li F, Butler KV, Kaniskan HU, Speed BA, Dela Seña C, et al. (2016). A potent, selective, and cell-active inhibitor of human type I protein arginine methyltransferases. ACS Chem. Biol. 11, 772–781.
    1. Fang J, Ye Z, Gu F, Yan M, Lin Q, Lin J, Wang Z, Xu Y, and Wang Y (2018). DUSP1 enhances the chemoresistance of gallbladder cancer via the modulation of the p38 pathway and DNA damage/repair system. Oncol. Lett. 16, 1869–1875.
    1. Fong JY, Pignata L, Goy PA, Kawabata KC, Lee SC, Koh CM, Musiani D, Massignani E, Kotini AG, Penson A, et al. (2019). Therapeutic targeting of RNA splicing catalysis through inhibition of protein arginine methylation. Cancer Cell 36, 194–209.e9.
    1. Garcia-Manero G, Khoury HJ, Jabbour E, Lancet J, Winski SL, Cable L, Rush S, Maloney L, Hogeland G, Ptaszynski M, et al. (2015). A phase I study of oral ARRY-614, a p38 MAPK/Tie2 dual inhibitor, in patients with low or intermediate-1 risk myelodysplastic syndromes. Clin. Cancer Res. 21, 985–994.
    1. Geest CR, and Coffer PJ (2009). MAPK signaling pathways in the regulation of hematopoiesis. J. Leukoc. Biol. 86, 237–250.
    1. Gerstung M, Pellagatti A, Malcovati L, Giagounidis A, Porta MG, Jädersten M, Dolatshad H, Verma A, Cross NC, Vyas P, et al. (2015). Combining gene mutation with gene expression data improves outcome prediction in myelodysplastic syndromes. Nat. Commun. 6, 5901.
    1. Greenberg P, Cox C, LeBeau MM, Fenaux P, Morel P, Sanz G, Sanz M, Vallespi T, Hamblin T, Oscier D, et al. (1997). International scoring system for evaluating prognosis in myelodysplastic syndromes. Blood 89, 2079– 2088.
    1. Guan KL, and Butch E (1995). Isolation and characterization of a novel dual specific phosphatase, HVH2, which selectively dephosphorylates the mitogen-activated protein kinase. J. Biol. Chem. 270, 7197–7203.
    1. Guccione E, and Richard S (2019). The regulation, functions and clinical relevance of arginine methylation. Nat. Rev. Mol. Cell Biol. 20, 642–657.
    1. Guo A, Gu H, Zhou J, Mulhern D, Wang Y, Lee KA, Yang V, Aguiar M, Kornhauser J, Jia X, et al. (2014a). Immunoaffinity enrichment and mass spectrometry analysis of protein methylation. Mol. Cell. Proteomics 13, 372–387.
    1. Guo H, Wang R, Zheng W, Chen Y, Blum G, Deng H, and Luo M (2014b). Profiling substrates of protein arginine N-methyltransferase 3 with S-adenosyl-L-methionine analogues. ACS Chem. Biol. 9, 476–484.
    1. Hayer A, Shao L, Chung M, Joubert LM, Yang HW, Tsai FC, Bisaria A, Betzig E, and Meyer T (2016). Engulfed cadherin fingers are polarized junctional structures between collectively migrating endothelial cells. Nat. Cell Biol. 18, 1311–1323.
    1. He X, Zhu Y, Lin YC, Li M, Du J, Dong H, Sun J, Zhu L, Wang H,Ding Z, et al. (2019). PRMT1-mediated FLT3 arginine methylation promotes maintenance of FLT3-ITD+ acute myeloid leukemia. Blood 134, 548–560.
    1. Herrera R, Hubbell S, Decker S, and Petruzzelli L (1998). A role for the MEK/MAPK pathway in PMA-induced cell cycle arrest: Modulation of megakaryocytic differentiation of K562 cells. Exp. Cell Res. 238, 407–414.
    1. Hofweber M, Hutten S, Bourgeois B, Spreitzer E, Niedner-Boblenz A, Schifferer M, Ruepp MD, Simons M, Niessing D, Madl T, and Dormann D (2018). Phase separation of FUS is suppressed by its nuclear import receptor and arginine methylation. Cell 173, 706–719.e13.
    1. Hu P, Nebreda AR, Hanenberg H, Kinnebrew GH, Ivan M, Yoder MC, Filippi MD, Broxmeyer HE, and Kapur R (2018). P38α/JNK signaling restrains erythropoiesis by suppressing Ezh2-mediated epigenetic silencing of Bim. Nat. Commun. 9, 3518.
    1. Hua WK, Chang YI, Yao CL, Hwang SM, Chang CY, and Lin WJ (2013). Protein arginine methyltransferase 1 interacts with and activates p38α to facilitate erythroid differentiation. PLoS ONE 8, e56715.
    1. Ibanez G, Shum D, Blum G, Bhinder B, Radu C, Antczak C, LUo M, and Djaballah H (2012). A high throughput scintillation proximity imaging assay for protein methyltransferases. Combinatorial chemistry & high throughput screening 15, 359–371.
    1. Ito K, Hirao A, Arai F, Takubo K, Matsuoka S, Miyamoto K, Ohmura M, Naka K, Hosokawa K, Ikeda Y, and Suda T (2006). Reactive oxygen species act through p38 MAPK to limit the lifespan of hematopoietic stem cells. Nat. Med. 12, 446–451.
    1. Jin S, Su H, Tran NT, Song J, Lu SS, Li Y, Huang S, Abdel-Wahab O, Liu Y, and Zhao X (2017). Splicing factor SF3B1K700E mutant dysregulates erythroid differentiation via aberrant alternative splicing of transcription factor TAL1. PLoS ONE 12, e0175523.
    1. Karigane D, Kobayashi H, Morikawa T, Ootomo Y, Sakai M, Nagamatsu G, Kubota Y, Goda N, Matsumoto M, Nishimura EK, et al. (2016). p38α activates purine metabolism to initiate hematopoietic stem/progenitor cell cycling in response to stress. Cell Stem Cell 19, 192–204.
    1. Kesarwani M, Kincaid Z, Gomaa A, Huber E, Rohrabaugh S, Siddiqui Z, Bouso MF, Latif T, Xu M, Komurov K, et al. (2017). Targeting c-FOS and DUSP1 abrogates intrinsic resistance to tyrosine-kinase inhibitor therapy in BCR-ABL-induced leukemia. Nat. Med. 23, 472–482.
    1. Keyse SM (2008). Dual-specificity MAP kinase phosphatases (MKPs) and cancer. Cancer Metastasis Rev. 27, 253–261.
    1. Klein AM, Zaganjor E, and Cobb MH (2013). Chromatin-tethered MAPKs. Curr. Opin. Cell Biol. 25, 272–277.
    1. Larsen SC, Sylvestersen KB, Mund A, Lyon D, Mullari M, Madsen MV, Daniel JA, Jensen LJ, and Nielsen ML (2016). Proteome-wide analysis of arginine monomethylation reveals widespread occurrence in human cells. Sci. Signal. 9, rs9.
    1. Lefrançais E, Ortiz-Muñoz G, Caudrillier A, Mallavia B, Liu F, Sayah DM, Thornton EE, Headley MB, David T, Coughlin SR, et al. (2017). The lung is a site of platelet biogenesis and a reservoir for haematopoietic progenitors. Nature 544, 105–109.
    1. Lim CP, Jain N, and Cao X (1998). Stress-induced immediate-early gene, egr-1, involves activation of p38/JNK1. Oncogene 16, 2915–2926.
    1. Liu ZJ, and Sola-Visner M (2011). Neonatal and adult megakaryopoiesis. Curr. Opin. Hematol. 18, 330–337.
    1. Liu M, Hsu J, Chan C, Li Z, and Zhou Q (2012). The ubiquitin ligase Siah1 controls ELL2 stability and formation of super elongation complexes to modulate gene transcription. Mol. Cell 46, 325–334.
    1. Lu YC, Sanada C, Xavier-Ferrucio J, Wang L, Zhang PX, Grimes HL, Venkatasubramanian M, Chetal K, Aronow B, Salomonis N, and Krause DS. (2018). The molecular signature of megakaryocyte-erythroid progenitors reveals a role for the cell cycle in fate specification. Cell Rep. 25, 3229.
    1. Luo M (2012). Current chemical biology approaches to interrogate protein methyltransferases. ACS Chem. Biol. 7, 443–463.
    1. Luo M (2018). Chemical and biochemical perspectives of protein lysine methylation. Chem. Rev. 118, 6656–6705.
    1. Mathioudaki K, Scorilas A, Ardavanis A, Lymberi P, Tsiambas E, Devetzi M, Apostolaki A, and Talieri M (2011). Clinical evaluation of PRMT1 gene expression in breast cancer. Tumour Biol. 32, 575–582.
    1. Mazharian A, Watson SP, and Séverin S (2009). Critical role for ERK1/2 in bone marrow and fetal liver-derived primary megakaryocyte differentiation, motility, and proplatelet formation. Exp. Hematol. 37, 1238–1249.e5.
    1. Miyazaki R, Ogata H, and Kobayashi Y (2001). Requirement of thrombopoietin-induced activation of ERK for megakaryocyte differentiation and of p38 for erythroid differentiation. Ann. Hematol. 80, 284–291.
    1. Mootha VK, Lindgren CM, Eriksson KF, Subramanian A, Sihag S, Lehar J, Puigserver P, Carlsson E, Ridderstråle M, Laurila E, et al. (2003). PGC-1α-responsive genes involved in oxidative phosphorylation are coordinately downregulated in human diabetes. Nat. Genet. 34, 267–273.
    1. Naeem H, Cheng D, Zhao Q, Underhill C, Tini M, Bedford MT, and Torchia J (2007). The activity and stability of the transcriptional coactivator p/CIP/SRC-3 are regulated by CARM1-dependent methylation. Mol. Cell. Biol. 27, 120–134.
    1. Nagata Y, Takahashi N, Davis RJ, and Todokoro K (1998). Activation of p38 MAP kinase and JNK but not ERK is required for erythropoietin-induced erythroid differentiation. Blood 92, 1859–1869.
    1. Navas TA, Mohindru M, Estes M, Ma JY, Sokol L, Pahanish P, Parmar S, Haghnazari E, Zhou L, Collins R, et al. (2006). Inhibition of overactivated p38 MAPK can restore hematopoiesis in myelodysplastic syndrome progenitors. Blood 108, 4170–4177.
    1. Nimer SD (2008). Myelodysplastic syndromes. Blood 111, 4841–4851.
    1. Norman TM, Horlbeck MA, Replogle JM, Ge AY, Xu A, Jost M, Gilbert LA, and Weissman JS (2019). Exploring genetic interaction manifolds constructed from rich single-cell phenotypes. Science 365, 786–793.
    1. Owens DM, and Keyse SM (2007). Differential regulation of MAP kinase signalling by dual-specificity protein phosphatases. Oncogene 26, 3203–3213.
    1. Patterson KI, Brummer T, O’Brien PM, and Daly RJ (2009). Dual-specificity phosphatases: Critical regulators with diverse cellular targets. Biochem. J. 418, 475–489.
    1. Pellagatti A, Cazzola M, Giagounidis A, Perry J, Malcovati L, Della Porta MG, Jädersten M, Killick S, Verma A, Norbury CJ, et al. (2010). Deregulated gene expression pathways in myelodysplastic syndrome hematopoietic stem cells. Leukemia 24, 756–764.
    1. Pellin D, Loperfido M, Baricordi C, Wolock SL, Montepeloso A, Weinberg OK, Biffi A, Klein AM, and Biasco L (2019). A comprehensive single cell transcriptional landscape of human hematopoietic progenitors. Nat. Commun. 10, 2395.
    1. Pronk CJ, Rossi DJ, Månsson R, Attema JL, Norddahl GL, Chan CK, Sigvardsson M, Weissman IL, and Bryder D (2007). Elucidation of the phenotypic, functional, and molecular topography of a myeloerythroid progenitor cell hierarchy. Cell Stem Cell 1, 428–442.
    1. Psaila B, and Mead AJ (2019). Single-cell approaches reveal novel cellular pathways for megakaryocyte and erythroid differentiation. Blood 133, 1427–1435.
    1. Qamar S, Wang G, Randle SJ, Ruggeri FS, Varela JA, Lin JQ, Phillips EC, Miyashita A, Williams D, Ströhl F, et al. (2018). FUS phase separation is modulated by a molecular chaperone and methylation of arginine cation-π interactions. Cell 173, 720–734.e15.
    1. Ramsey SD, McCune JS, Blough DK, McDermott CL, Beck SJ, López JA, and Deeg HJ (2012). Patterns of blood product use among patients with myelodysplastic syndrome. Vox Sang. 102, 331–337.
    1. Reya T, Duncan AW, Ailles L, Domen J, Scherer DC, Willert K, Hintz L, Nusse R, and Weissman IL (2003). A role for Wnt signalling in self-renewal of haematopoietic stem cells. Nature 423, 409–414.
    1. Robinson CJ, Sloss CM, and Plevin R (2001). Inactivation of JNK activity by mitogen-activated protein kinase phosphatase-2 in EAhy926 endothelial cells is dependent upon agonist-specific JNK translocation to the nucleus. Cell. Signal. 13, 29–41.
    1. Rodriguez-Fraticelli AE, Wolock SL, Weinreb CS, Panero R, Patel SH, Jankovic M, Sun J, Calogero RA, Klein AM, and Camargo FD (2018). Clonal analysis of lineage fate in native haematopoiesis. Nature 553, 212–216.
    1. Sakamaki J, Daitoku H, Ueno K, Hagiwara A, Yamagata K, and Fukamizu A (2011). Arginine methylation of BCL-2 antagonist of cell death (BAD) counteracts its phosphorylation and inactivation by Akt. Proc. Natl. Acad. Sci. USA 108, 6085–6090.
    1. Sanada C, Xavier-Ferrucio J, Lu YC, Min E, Zhang PX, Zou S, Kang E, Zhang M, Zerafati G, Gallagher PG, and Krause DS (2016). Adult human megakaryocyte-erythroid progenitors are in the CD34+CD38mid fraction. Blood 128, 923–933.
    1. Sanjuan-Pla A, Macaulay IC, Jensen CT, Woll PS, Luis TC, Mead A, Moore S, Carella C, Matsuoka S, Bouriez Jones T, et al. (2013). Platelet-biased stem cells reside at the apex of the haematopoietic stem-cell hierarchy. Nature 502, 232–236.
    1. Schinke C, Giricz O, Li W, Shastri A, Gordon S, Barreyro L, Bhagat T, Bhattacharyya S, Ramachandra N, Bartenstein M, et al. (2015). IL8-CXCR2 pathway inhibition as a therapeutic strategy against MDS and AML stem cells. Blood 125, 3144–3152.
    1. Seternes OM, Kidger AM, and Keyse SM (2019). Dual-specificity MAP kinase phosphatases in health and disease. Biochim. Biophys. Acta Mol. Cell Res. 1866, 124–143.
    1. Suganuma T, and Workman JL (2011). Signals and combinatorial functions of histone modifications. Annu. Rev. Biochem. 80, 473–499.
    1. Sun S, Wang W, Latchman Y, Gao D, Aronow B, and Reems JA (2013). Expression of plasma membrane receptor genes during megakaryocyte development. Physiol. Genomics 45, 217–227.
    1. Tusi BK, Wolock SL, Weinreb C, Hwang Y, Hidalgo D, Zilionis R, Waisman A, Huh JR, Klein AM, and Socolovsky M (2018). Population snapshots predict early haematopoietic and erythroid hierarchies. Nature 555, 54–60.
    1. Uddin S, Ah-Kang J, Ulaszek J, Mahmud D, and Wickrema A (2004). Differentiation stage-specific activation of p38 mitogen-activated protein kinase isoforms in primary human erythroid cells. Proc. Natl. Acad. Sci. USA 101, 147–152.
    1. van Dijk D, Sharma R, Nainys J, Yim K, Kathail P, Carr AJ, Burdziak C, Moon KR, Chaffer CL, Pattabiraman D, et al. (2018). Recovering gene interactions from single-cell data using data diffusion. Cell 174, 716–729.e27.
    1. Verma A, Deb DK, Sassano A, Uddin S, Varga J, Wickrema A, and Platanias LC (2002). Activation of the p38 mitogen-activated protein kinase mediates the suppressive effects of type I interferons and transforming growth factor-b on normal hematopoiesis. J. Biol. Chem. 277, 7726–7735.
    1. Wan W, You Z, Zhou L, Xu Y, Peng C, Zhou T, Yi C, Shi Y, and Liu W (2018). mTORC1-regulated and HUWE1-mediated WIPI2 degradation controls autophagy flux. Mol. Cell 72, 303–315.e6.
    1. Wang R, and Luo M (2013). A journey toward bioorthogonal profiling of protein methylation inside living cells. Curr. Opin. Chem. Biol. 17, 729–737.
    1. Wang Q, Miyakawa Y, Fox N, and Kaushansky K (2000). Interferon-α directly represses megakaryopoiesis by inhibiting thrombopoietin-induced signaling through induction of SOCS-1. Blood 96, 2093–2099.
    1. Wang R, Zheng W, Yu H, Deng H, and Luo M (2011). Labeling substrates of protein arginine methyltransferase with engineered enzymes and matched S-adenosyl-L-methionine analogues. J. Am. Chem. Soc. 133, 7648–7651.
    1. Wang R, Islam K, Liu Y, Zheng W, Tang H, Lailler N, Blum G, Deng H, and Luo M (2013). Profiling genome-wide chromatin methylation with engineered posttranslation apparatus within living cells. J. Am. Chem. Soc. 135, 1048–1056.
    1. Wang R, Zheng W, and Luo M (2014). A sensitive mass spectrum assay to characterize engineered methionine adenosyltransferases with S-alkyl methionine analogues as substrates. Anal. Biochem. 450, 11–19.
    1. Wang Q, Reszka-Blanco N, Cheng L, Li G, Zhang L, and Su L (2018). p38 MAPK is critical for nuclear translocation of IRF-7 during CpG-induced type I IFN expression in human plasmacytoid dendritic cells. J. Immunol. 200 (Suppl), 109.106.
    1. Weinreb C, Wolock S, and Klein AM (2018). SPRING: A kinetic interface for visualizing high dimensional single-cell expression data. Bioinformatics 34, 1246–1248.
    1. Weinreb C, Rodriguez-Fraticelli A, Camargo FD, and Klein AM (2020). Lineage tracing on transcriptional landscapes links state to fate during differentiation. Science 367, eaaw3381.
    1. Whalen AM, Galasinski SC, Shapiro PS, Nahreini TS, and Ahn NG (1997). Megakaryocytic differentiation induced by constitutive activation of mitogen-activated protein kinase kinase. Mol. Cell. Biol. 17, 1947–1958.
    1. Witt O, Sand K, and Pekrun A (2000). Butyrate-induced erythroid differentiation of human K562 leukemia cells involves inhibition of ERK and activation of p38 MAP kinase pathways. Blood 95, 2391–2396.
    1. Xavier-Ferrucio J, and Krause DS (2018). Concise review: Bipotent megakaryocytic-erythroid progenitors: Concepts and controversies. Stem Cells 36, 1138–1145.
    1. Yamagata K, Daitoku H, Takahashi Y, Namiki K, Hisatake K, Kako K, Mukai H, Kasuya Y, and Fukamizu A (2008). Arginine methylation of FOXO transcription factors inhibits their phosphorylation by Akt. Mol. Cell 32, 221–231.
    1. Yamamoto R, Morita Y, Ooehara J, Hamanaka S, Onodera M, Rudolph KL, Ema H, and Nakauchi H (2013). Clonal analysis unveils self-renewing lineage-restricted progenitors generated directly from hematopoietic stem cells. Cell 154, 1112–1126.
    1. Yu ZH, and Zhang ZY (2018). Regulatory mechanisms and novel therapeutic targeting strategies for protein tyrosine phosphatases. Chem. Rev. 118, 1069–1091.
    1. Zhang L, Tran NT, Su H, Wang R, Lu Y, Tang H, Aoyagi S, Guo A, Khodadadi-Jamayran A, Zhou D, et al. (2015). Cross-talk between PRMT1-mediated methylation and ubiquitylation on RBM15 controls RNA splicing. eLife 4, e07938.
    1. Zhao X, Jankovic V, Gural A, Huang G, Pardanani A, Menendez S, Zhang J, Dunne R, Xiao A, Erdjument-Bromage H, et al. (2008). Methylation of RUNX1 by PRMT1 abrogates SIN3A binding and potentiates its transcriptional activity. Genes Dev. 22, 640–653.
    1. Zhu L, He X, Dong H, Sun J, Wang H, Zhu Y, Huang F, Zou J, Chen Z, Zhao X, and Li L (2019). Protein arginine methyltransferase 1 is required for maintenance of normal adult hematopoiesis. Int. J. Biol. Sci. 15, 2763–2773.

Source: PubMed

3
Abonnere