15
Fafilek et al., Sci. Signal. 11, eaap8608 (2018) 18 September 2018 SCIENCE SIGNALING | RESEARCH ARTICLE 1 of 14 CELL BIOLOGY The inositol phosphatase SHIP2 enables sustained ERK activation downstream of FGF receptors by recruiting Src kinases Bohumil Fafilek 1,2 *, Lukas Balek 1 *, Michaela Kunova Bosakova 1,2 , Miroslav Varecha 1,2 , Alexandru Nita 1 , Tomas Gregor 3 , Iva Gudernova 1 , Jitka Krenova 1 , Somadri Ghosh 4 , Martin Piskacek 5 , Lucie Jonatova 1 , Nicole H. Cernohorsky 1 , Jennifer T. Zieba 6 , Michal Kostas 7,8 , Ellen Margrethe Haugsten 7,8 , Jørgen Wesche 7,8 , Christophe Erneux 4 , Lukas Trantirek 3 , Deborah Krakow 6,9,10 , Pavel Krejci 1,2,11† Sustained activation of extracellular signal–regulated kinase (ERK) drives pathologies caused by mutations in fibroblast growth factor receptors (FGFRs). We previously identified the inositol phosphatase SHIP2 (also known as INPPL1) as an FGFR-interacting protein and a target of the tyrosine kinase activities of FGFR1, FGFR3, and FGFR4. We report that loss of SHIP2 converted FGF-mediated sustained ERK activation into a transient signal and rescued cell phenotypes triggered by pathologic FGFR-ERK signaling. Mutant forms of SHIP2 lacking phospho- inositide phosphatase activity still associated with FGFRs and did not prevent FGF-induced sustained ERK activation, demonstrating that the adaptor rather than the catalytic activity of SHIP2 was required. SHIP2 recruited Src family kinases to the FGFRs, which promoted FGFR-mediated phosphorylation and assembly of protein complexes that relayed signaling to ERK. SHIP2 interacted with FGFRs, was phosphorylated by active FGFRs, and promoted FGFR-ERK signaling at the level of phosphorylation of the adaptor FRS2 and recruitment of the tyrosine phosphatase PTPN11. Thus, SHIP2 is an essential component of canonical FGF-FGFR signal transduction and a potential therapeutic target in FGFR-related disorders. INTRODUCTION Maintenance of tissue homeostasis depends on complex intercellular signaling networks that govern basic cell functions. The fibroblast growth factor (FGF) signaling network is a major contributor to such cell-to-cell communications. The four mammalian FGF receptors (FGFR1 to FGFR4) are receptor tyrosine kinases (RTKs) that respond to extracellular signals delivered by at least 18 FGF ligands during development, postnatal life, and disease. The importance of FGFR signaling is emphasized by their role in pathological functions. Various disorders arise from mutations, gene fusions, increased copy numbers, and other lesions affecting FGFR genes. These include cancer, developmental defects, and bone and skin disorders. Acti- vating mutations in FGFR3 alone are associated with five severe skeletal dysplasias, nine types of cancer, and two skin syndromes (14). Signaling through the extracellular signal–regulated kinase (ERK), a mitogen-activated protein (MAP) kinase, is implicated in most cellular phenotypes affected by FGF signaling. In human embryonic stem cells (hESCs) and induced pluripotent cells, FGFR-mediated ERK activation is critical for maintenance of the undifferentiated phenotype (57). Activation of ERK is the predominant mechanism through which FGFRs trigger cell proliferation in fibroblasts, endo- thelial cells, myoblasts, mesenchymal stem cells, neural progenitors, and lung and lens epithelial cells (1). However, aberrant sustained ERK activation mediates FGFR oncogenic signaling in FGFR4-driven prostate cancer, BaF3 lymphocytes and leukemic KG1 cells that express cancer-associated FGFR3 and FGFR1 fusion proteins that result from chromosomal translocations, respectively, and KMS11 multiple myeloma cells overexpressing FGFR3 as a result of a chromo- somal translocation (811). ERK drives much of the pathology underlying FGFR3-related skeletal dysplasias that include achondro- plasia (ACH) and thanatophoric dysplasia (TD). Sustained ERK activation due to heterozygosity for FGFR3-activating mutations that cause ACH and TD results in decreased chondrocyte proliferation, loss of extracellular matrix (ECM) in the growth plate cartilage, and altered chondrocyte differentiation (12, 13). Similarly, increased ERK signaling causes premature fusion of synchondroses in the developing vertebrae, which results in a narrowing of the spinal canal at the foramen magnum that contributes to the skeletal complications in ACH (14). Expression of constitutively active MEK1, the MAP kinase kinase upstream of ERK, produces similar phenotypic con- sequences as ERK overactivation, leading to an ACH-like dwarfism in mice (15). FGFRs stimulate ERK signaling through adaptor-mediated translocation of the RAS guanine nucleotide exchange factor SOS1 to the cell membrane. FGFR substrate 2 (FRS2) is a major adaptor involved in this process. FGFRs phosphorylate FRS2 on at least six tyrosine residues that serve as binding sites for SOS1 complexed 1 Department of Biology, Masaryk University, 62500 Brno, Czech Republic. 2 Interna- tional Clinical Research Center, St. Anne’s University Hospital, 65691 Brno, Czech Republic. 3 Central European Institute of Technology, Masaryk University, 62500 Brno, Czech Republic. 4 Institut de Recherche Interdisciplinaire en Biologie Humaine et moléculaire, Université Libre de Bruxelles, 1070 Bruxelles, Belgium. 5 Department of Pathological Physiology, Faculty of Medicine, Masaryk University, 62500 Brno, Czech Republic. 6 Department of Orthopedic Surgery, University of California Los Angeles, CA 90095, USA. 7 Department of Tumor Biology, Institute for Cancer Re- search, Norwegian Radium Hospital, 0379 Oslo, Norway. 8 Institute of Clinical Medicine, Faculty of Medicine, University of Oslo, 0379 Oslo, Norway. 9 Depart- ment of Human Genetics, University of California Los Angeles, CA 90095, USA. 10 Department of Obstetrics and Gynecology, David Geffen School of Medicine, University of California Los Angeles, CA 90095, USA. 11 Institute of Animal Physiol- ogy and Genetics of the Czech Academy of Sciences, 60200 Brno, Czech Republic. *These authors contributed equally to this work. †Corresponding author. Email: [email protected] Copyright © 2018 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original U.S. Government Works on February 21, 2020 http://stke.sciencemag.org/ Downloaded from

CELL BIOLOGY Copyright © 2018 The inositol phosphatase ... · Fafilek et al., Sci. Signal. 11, eaap8608 2018 18 September 2018 SCIENCE SIGNALING| RESEARCH ARTICLE 1 of 14 CELL BIOLOGY

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: CELL BIOLOGY Copyright © 2018 The inositol phosphatase ... · Fafilek et al., Sci. Signal. 11, eaap8608 2018 18 September 2018 SCIENCE SIGNALING| RESEARCH ARTICLE 1 of 14 CELL BIOLOGY

Fafilek et al., Sci. Signal. 11, eaap8608 (2018) 18 September 2018

S C I E N C E S I G N A L I N G | R E S E A R C H A R T I C L E

1 of 14

C E L L B I O L O G Y

The inositol phosphatase SHIP2 enables sustained ERK activation downstream of FGF receptors by recruiting Src kinasesBohumil Fafilek1,2*, Lukas Balek1*, Michaela Kunova Bosakova1,2, Miroslav Varecha1,2, Alexandru Nita1, Tomas Gregor3, Iva Gudernova1, Jitka Krenova1, Somadri Ghosh4, Martin Piskacek5, Lucie Jonatova1, Nicole H. Cernohorsky1, Jennifer T. Zieba6, Michal Kostas7,8, Ellen Margrethe Haugsten7,8, Jørgen Wesche7,8, Christophe Erneux4, Lukas Trantirek3, Deborah Krakow6,9,10, Pavel Krejci1,2,11†

Sustained activation of extracellular signal–regulated kinase (ERK) drives pathologies caused by mutations in fibroblast growth factor receptors (FGFRs). We previously identified the inositol phosphatase SHIP2 (also known as INPPL1) as an FGFR-interacting protein and a target of the tyrosine kinase activities of FGFR1, FGFR3, and FGFR4. We report that loss of SHIP2 converted FGF-mediated sustained ERK activation into a transient signal and rescued cell phenotypes triggered by pathologic FGFR-ERK signaling. Mutant forms of SHIP2 lacking phospho-inositide phosphatase activity still associated with FGFRs and did not prevent FGF-induced sustained ERK activation, demonstrating that the adaptor rather than the catalytic activity of SHIP2 was required. SHIP2 recruited Src family kinases to the FGFRs, which promoted FGFR-mediated phosphorylation and assembly of protein complexes that relayed signaling to ERK. SHIP2 interacted with FGFRs, was phosphorylated by active FGFRs, and promoted FGFR-ERK signaling at the level of phosphorylation of the adaptor FRS2 and recruitment of the tyrosine phosphatase PTPN11. Thus, SHIP2 is an essential component of canonical FGF-FGFR signal transduction and a potential therapeutic target in FGFR-related disorders.

INTRODUCTIONMaintenance of tissue homeostasis depends on complex intercellular signaling networks that govern basic cell functions. The fibroblast growth factor (FGF) signaling network is a major contributor to such cell-to-cell communications. The four mammalian FGF receptors (FGFR1 to FGFR4) are receptor tyrosine kinases (RTKs) that respond to extracellular signals delivered by at least 18 FGF ligands during development, postnatal life, and disease. The importance of FGFR signaling is emphasized by their role in pathological functions. Various disorders arise from mutations, gene fusions, increased copy numbers, and other lesions affecting FGFR genes. These include cancer, developmental defects, and bone and skin disorders. Acti-vating mutations in FGFR3 alone are associated with five severe skeletal dysplasias, nine types of cancer, and two skin syndromes (1–4).

Signaling through the extracellular signal–regulated kinase (ERK), a mitogen-activated protein (MAP) kinase, is implicated in most

cellular phenotypes affected by FGF signaling. In human embryonic stem cells (hESCs) and induced pluripotent cells, FGFR-mediated ERK activation is critical for maintenance of the undifferentiated phenotype (5–7). Activation of ERK is the predominant mechanism through which FGFRs trigger cell proliferation in fibroblasts, endo-thelial cells, myoblasts, mesenchymal stem cells, neural progenitors, and lung and lens epithelial cells (1). However, aberrant sustained ERK activation mediates FGFR oncogenic signaling in FGFR4-driven prostate cancer, BaF3 lymphocytes and leukemic KG1 cells that express cancer-associated FGFR3 and FGFR1 fusion proteins that result from chromosomal translocations, respectively, and KMS11 multiple myeloma cells overexpressing FGFR3 as a result of a chromo-somal translocation (8–11). ERK drives much of the pathology underlying FGFR3-related skeletal dysplasias that include achondro-plasia (ACH) and thanatophoric dysplasia (TD). Sustained ERK activation due to heterozygosity for FGFR3-activating mutations that cause ACH and TD results in decreased chondrocyte proliferation, loss of extracellular matrix (ECM) in the growth plate cartilage, and altered chondrocyte differentiation (12, 13). Similarly, increased ERK signaling causes premature fusion of synchondroses in the developing vertebrae, which results in a narrowing of the spinal canal at the foramen magnum that contributes to the skeletal complications in ACH (14). Expression of constitutively active MEK1, the MAP kinase kinase upstream of ERK, produces similar phenotypic con-sequences as ERK overactivation, leading to an ACH-like dwarfism in mice (15).

FGFRs stimulate ERK signaling through adaptor-mediated translocation of the RAS guanine nucleotide exchange factor SOS1 to the cell membrane. FGFR substrate 2 (FRS2) is a major adaptor involved in this process. FGFRs phosphorylate FRS2 on at least six tyrosine residues that serve as binding sites for SOS1 complexed

1Department of Biology, Masaryk University, 62500 Brno, Czech Republic. 2Interna-tional Clinical Research Center, St. Anne’s University Hospital, 65691 Brno, Czech Republic. 3Central European Institute of Technology, Masaryk University, 62500 Brno, Czech Republic. 4Institut de Recherche Interdisciplinaire en Biologie Humaine et moléculaire, Université Libre de Bruxelles, 1070 Bruxelles, Belgium. 5Department of Pathological Physiology, Faculty of Medicine, Masaryk University, 62500 Brno, Czech Republic. 6Department of Orthopedic Surgery, University of California Los Angeles, CA 90095, USA. 7Department of Tumor Biology, Institute for Cancer Re-search, Norwegian Radium Hospital, 0379 Oslo, Norway. 8Institute of Clinical Medicine, Faculty of Medicine, University of Oslo, 0379 Oslo, Norway. 9Depart-ment of Human Genetics, University of California Los Angeles, CA 90095, USA. 10Department of Obstetrics and Gynecology, David Geffen School of Medicine, University of California Los Angeles, CA 90095, USA. 11Institute of Animal Physiol-ogy and Genetics of the Czech Academy of Sciences, 60200 Brno, Czech Republic.*These authors contributed equally to this work.†Corresponding author. Email: [email protected]

Copyright © 2018 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original U.S. Government Works

on February 21, 2020

http://stke.sciencemag.org/

Dow

nloaded from

Page 2: CELL BIOLOGY Copyright © 2018 The inositol phosphatase ... · Fafilek et al., Sci. Signal. 11, eaap8608 2018 18 September 2018 SCIENCE SIGNALING| RESEARCH ARTICLE 1 of 14 CELL BIOLOGY

Fafilek et al., Sci. Signal. 11, eaap8608 (2018) 18 September 2018

S C I E N C E S I G N A L I N G | R E S E A R C H A R T I C L E

2 of 14

with growth factor receptor–bound protein 2 (GRB2), directing a substantial amount of SOS1 to the cell membrane where it activates RAS (16, 17).

Sustained ERK activation downstream of FGFR signaling is normal in some contexts but pathogenic when it occurs inappropriately. In chondrocytes and hESCs, FGFR activation leads to strong FRS2 tyrosine phosphorylation and corresponding sustained ERK activa-tion lasting for more than 8 hours (18). The molecular processes underlying sustained ERK activity in FGFR signaling are poorly characterized. Here, we report that FGFRs interacted with and phosphorylated SRC homology 2 (SH2) domain–containing inositol 5-phosphatase 2 [SHIP2; also known as inositol polyphosphate phosphatase–like 1 (INPPL1)], which, in turn, recruited Src family kinases (SFKs) to the FGFR signaling complex, where they promoted FGFR-mediated phosphorylation of the adaptors FRS2 and GRB2-associated binding protein 1 (GAB1) to relay the signal to the ERK pathway. Loss of SHIP2 converted sustained FGF-mediated ERK activation into a transient signal, demonstrating that SHIP2 is an essential component of FGF signaling and necessary for the maintenance of FGF-mediated ERK activation.

RESULTSFGFRs interact with SHIP2 and target SHIP2 to focal adhesionsIn a previous study, we used proteomics to identify proteins that were phosphorylated upon activation of endogenous Fgfr2 and Fgfr3 (19) signaling in cultured rat chondrosarcoma (RCS) cells. As detailed previously, the tyrosine phosphorylated proteins were purified from FGF2-treated cells by immunoprecipitation (IP) with an antibody that recognizes phosphorylated tyrosine residues and subjected to tandem mass spectrometry (19). The inositol phos-phatase SHIP2 was among the most frequently identified proteins phosphorylated at tyrosine upon FGFR activation, being found in four of the six experiments we performed (19).

Both wild-type FGFR3 and the activating mutant FGFR3- K650M, which is associated with TD (20), coimmunoprecipitated with human SHIP2 when expressed in 293T cells (Fig. 1A). FGFR3 induced SHIP2 phosphorylation at Tyr986, Tyr987, and Tyr1135 in 293T cells (Fig. 1A) and in cell-free kinase assays using recombinant FGFR3 and recom-binant SHIP2 (recSHIP2; Fig. 1B). Addition of recSHIP2 did not affect the activity of FGFR3 in a kinase assay, as demonstrated by no change in FGFR3 autophosphorylation in the presence of SHIP2 (Fig. 1C). Similarly, FGFR3 did not alter the phosphoinositide phos-phatase (PS) activity of SHIP2, as demonstrated by a cell-free SHIP2 activity assay based on colorimetric detection of phosphate released during the hydrolysis of phosphatidylinositol (3,4,5)-trisphosphate [PtdIns(3,4,5)P3] to phosphatidylinositol (3,4,5)-bisphosphate [PtdIns(3,4)P2; Fig. 1D]. These experiments established that FGFR3 interacted with SHIP2 and acted as a SHIP2 kinase. The experiments also show that the FGFR3-SHIP2 interaction did not affect the catalytic activity of either protein.

SHIP2 contains three distinguished domains according to struc-tural data available in the Protein Data Bank (PDB) database (www.pdb.org): an N-terminal SH2 domain (PDB: 2MK2), a central PS domain (PDB: 4A9C), and a C-terminal sterile alpha motif (SAM) domain (PDB: 2K4P). In addition, we identified two proline-rich (PR) domains in SHIP2, located at residues 123 to 160 (PR1) and residues 935 to 1105 (PR2) in human SHIP2 (Fig. 2A). We generated

a series of C-terminally V5-tagged truncated human SHIP2 variants, coexpressed each with wild-type FGFR3 in 293T cells, and deter-mined whether each SHIP2 variant coimmunoprecipitated with FGFR3. Deletion of the PS or PR1 domain had no effect on SHIP2 coimmunoprecipitation with FGFR3 (Fig. 2, B and C). In contrast, deletion of the SH2 or SAM domain substantially impaired FGFR3- SHIP2 association [Fig. 2, D and E (arrows)], suggesting a bipartite SHIP2-FGFR3 interaction involving both the SH2 and SAM domains of SHIP2. Wild-type SHIP2 did not coimmunoprecipitate with the catalytically inactive K508M mutant form of FGFR3 (Fig. 2F), sug-gesting that SHIP2 associates preferentially with active FGFR3.

Having found that SHIP2 interacted with and was phosphorylated by FGFR3, we asked whether SHIP2 also associated with other FGFRs. In 293T cells, SHIP2 coimmunoprecipitated with FGFR1 and FGFR4 but not with FGFR2 (Fig. 3A). We next used osteosarcoma U2OS cells stably expressing a C-terminally green fluorescent protein (GFP)–tagged FGFR1 to probe the association of endogenous SHIP2 with FGFR1. Proximity ligation assays (PLAs) using primary anti-bodies recognizing SHIP2 and GFP revealed that endogenous SHIP2 colocalized with transgenically expressed FGFR1 (Fig. 3, B and C). We obtained similar data using PLA with primary antibodies recog-nizing FGFR1 and SHIP2 (Fig. 3, B and D). We detected no sub-stantial PLA signal in isogenic negative controls for both experiments: human osteosarcoma U2OS cells stably expressing FGFR1–BirA (FGFR1 fused to biotin ligase)–HA (hemagglutinin tag) (for GFP: SHIP2 PLA) and U2OS cells stably expressing FGFR4- GFP (for FGFR1:SHIP2 PLA).

Treatment of RCS cells with FGF2 triggered phosphorylation of endogenous Ship2 at Tyr986 and Tyr987 (Fig. 4A) and targeted Ship2 to the cell periphery, where it partially colocalized with the focal adhesion marker vinculin (Fig. 4, B and C). We observed similar translocation and association with peripheral focal adhesions for the signaling adaptor p130Cas (Fig. 4, D and E), which was previously found to be phosphorylated at multiple tyrosines in FGF2-treated RCS cells (19). Because Ship2 and p130Cas interacted with FGFR3 and are FGFR3 substrates, it is likely that they are phosphorylated by FGFR3 and transit to focal adhesions together. In support of this hypothesis, we previously reported that p130Cas and SHIP2 interact through the SH3 domain of p130Cas and localize to lamellipodia together (21).

SHIP2 is part of the integrin adhesome and is known to interact with various cytoskeletal proteins (22). Previous data obtained in human fibroblasts show that general cell migration is substantially reduced in SHIP2-null fibroblasts derived from opsismodysplasia (OPS) patients (23). To determine whether Ship2 plays a role in FGF-mediated cell migration, we used the CRISPR (clustered regular-ly interspaced short palindromic repeat)/Cas9 (CRISPR-associated) system to disrupt the Ship2 gene in RCS cells. We selected four RCS clones with the Ship2 locus targeted by CRISPR/Cas9 (Ship2Crispr cells), analyzed their genotypes by DNA sequencing, and determined whether Ship2 was present in these cells by Western blotting (clone names and genotypes: Ship2c−/−, Ship2a+/−, Ship2g−/−, and Ship2 f+/−; Fig. 4F). Using fibronectin-coated glass bottom chambers, we com-pared the cell migration velocity of wild-type and Ship2Crispr cells. In the presence of 10% serum, migration was inhibited in two Ship2Crispr cell lines compared to wild-type cells (Fig. 4G). In contrast, Ship2 deletion significantly increased migration in cells treated with FGF2 in the absence of serum (Fig. 4G). Therefore, Ship2 acts as a nega-tive regulator of FGF-mediated migration in RCS cells.

on February 21, 2020

http://stke.sciencemag.org/

Dow

nloaded from

Page 3: CELL BIOLOGY Copyright © 2018 The inositol phosphatase ... · Fafilek et al., Sci. Signal. 11, eaap8608 2018 18 September 2018 SCIENCE SIGNALING| RESEARCH ARTICLE 1 of 14 CELL BIOLOGY

Fafilek et al., Sci. Signal. 11, eaap8608 (2018) 18 September 2018

S C I E N C E S I G N A L I N G | R E S E A R C H A R T I C L E

3 of 14

Ship2 deletion rescues cellular phenotypes caused by FGFR activationRCS cells are a well-established chondrocyte model for studying pathologic increased FGFR3 signaling as seen in ACH or TD due to activating FGFR3 mutations. The cells respond to activation of endogenous FGFR signaling with complex changes in cell behavior including growth arrest, diminished production of collagen and pro-teoglycan ECM components, induction of premature senescence, and alteration in cell shape (12, 13, 24–26). FGF2 stimulation induced

less cell growth arrest in Ship2Crispr RCS cells compared to wild-type RCS cells (Fig. 5A). This phenotype was not due to clonal variation because three ran-domly selected RCS clones with wild-type Ship2 genotypes responded to FGF2 with growth arrest comparable to wild-type cells (Fig.  5, B and C). The Ship2Crispr cells also showed less FGF2- induced ECM loss, as determined by Western blotting for collagen 2 and Al-cian blue staining for sulfated proteo-glycans (Fig. 5, D and E, and fig. S1A). Finally, FGF-mediated induction of pre-mature senescence was also less pro-nounced in Ship2Crispr cells compared to wild-type controls, as demonstrated by Western blotting for the presence of the senescence marker caveolin 1 (Fig. 5D) (12). These experiments established that Ship2 deletion rendered RCS cells less responsive to FGF stimulation with re-spect to multiple FGF-mediated cellular phenotypes. In contrast, Ship2 deletion did not affect FGF-induced changes in cellular shape, manifested as cell flat-tening and enlargement (Fig. 5F and fig. S1B) due to cytoskeletal remodeling with abundant formation of actin stress fibers (Fig. 5G) (18). The translocation of Ship2 to peripheral focal adhesions in response to FGF2 (Fig. 4, B and C) was therefore not essential for FGF- mediated cell spreading and cytoskeletal remodeling.

Ship2 deletion converts FGF-mediated sustained ERK activation into a transient signalIt is possible that the lack of response to FGF2 in Ship2Crispr RCS cells may simply have been the result of reduced FGFR activity, but we found no differences be-tween the amounts of Fgfr2 or Fgfr3 in wild-type and Ship2Crispr cells (fig. S2, A and B). Because sustained Erk activation accounts for most of cellular phenotypes triggered by FGF in RCS cells (27), we examined the kinetics of Erk activity in Ship2Crispr cells. All four Ship2Crispr cell

lines showed markedly impaired FGF-mediated Erk activation, which progressed with time, resulting in diminished Erk signaling in cells treated with FGF2 for more than 4 hours (Fig. 6A and figs. S3, A to C, and S4, A to D). In contrast, wild-type cells maintained phos-phorylated Erk for at least 10 hours. Thus, the Ship2Crispr cells initial-ly responded to FGF-mediated Erk activation normally but failed to maintain Erk activity over prolonged periods of time. The defective maintenance of Erk activation accounts for the rescue of the FGF- mediated changes in cell behavior observed in Ship2Crispr cells.

A SHIP2-FLAG

FGFR3

Actin

FGFR3 IP V5

SHIP2

pSHIP2Y1135

pSHIP2Y986/7

SHIP2

FGFR3

WB SHIP2

FGFR3 IP FLAG

WB

FGFR3-WT-V5 FGFR3-K650M-V5

pFGFR3Y653/4

– – – + + + – – – + – + – – – + + – Input SHIP2

SHIP2-FLAG

FGFR3

SHIP2

pSHIP2

Kinase assay 40′ recSHIP2

FGFR3 ATP

+ – – + + + + + – + + + + – – + + +

+ SU5402 20′ – + +

+ – +

C

pFGFR3 (4G10)

SHIP2

Abs

orba

nce

(620

nm

)

B recSHIP2 – + + + FGFR3 + + – +

ATP + – + +

SHIP2

pSHIP2Y497

pSHIP2 (4G10)

pSHIP2Y1135

pSHIP2Y986/7

FGFR3

n = 3 D

FGFR3

n.s.

Fig. 1. FGFRs interact with SHIP2 and phosphorylate SHIP2 in cells. (A) 293T cells were transfected with V5-tagged wild-type (WT) FGFR3 or FGFR3-K650M together with FLAG-tagged SHIP2 and subjected to V5 or FLAG IP. Immunoprecipitates were separated by SDS–polyacrylamide gel electrophoresis (PAGE) and subjected to Western blotting (WB) with antibodies to detect FGFR3, FGFR3 phosphorylated on Tyr653 or Tyr654 (pFGFR3Y653/4), SHIP2, SHIP2-FLAG, and SHIP2 phosphorylated on Tyr986 or Tyr987 (pSHIP2Y986/7). Actin is a loading control. Note the SHIP2 phosphorylation in cells transfected with the activated FGFR3 mutant FGFR3-K650M. Immunoblotting for SHIP2 de-tects both endogenous and transgenically expressed SHIP2. (B) Cell-free kinase assays were carried out with purified recombinant FGFR3 and recSHIP2 proteins in the presence of adenosine 5′-triphosphate (ATP), and the reactions were separated by SDS-PAGE and subjected to Western blotting. The sample without ATP is a negative control. Phos-phorylation of SHIP2 (pSHIP2) was detected with 4G10, an antibody that recognizes all phosphotyrosine residues. (C) Cell-free kinase assays were repeated in the presence and absence of the FGFR inhibitor (SU5402) and immuno-blotted to show phosphorylated and nonphosphorylated forms of FGFR3 and SHIP2. Data in (A) to (C) are represent-ative of three independent experiments. (D) Cell-free colorimetric SHIP2 phosphatase activity assay measuring SHIP2- mediated hydrolysis of phosphoinositide PtdIns(3,4,5)P3 (PIP3) to PtdIns(3,4)P2 (PIP2). Purified, recSHIP2, and recombinant FGFR3 were added to reaction buffer in the presence of absence of PIP3 as indicated. Samples without ATP are negative controls for FGFR3 activity. Significance was assessed using Welch’s t test. Each column represents three independent experiments with SD indicated. For each experiment, the values were calculated as averages from two technical duplicates, each measured three times. n.s., not significant.

on February 21, 2020

http://stke.sciencemag.org/

Dow

nloaded from

Page 4: CELL BIOLOGY Copyright © 2018 The inositol phosphatase ... · Fafilek et al., Sci. Signal. 11, eaap8608 2018 18 September 2018 SCIENCE SIGNALING| RESEARCH ARTICLE 1 of 14 CELL BIOLOGY

Fafilek et al., Sci. Signal. 11, eaap8608 (2018) 18 September 2018

S C I E N C E S I G N A L I N G | R E S E A R C H A R T I C L E

4 of 14

We next asked whether the defect in FGF-mediated Erk activa-tion in Ship2 null cells could be restored by microinjection of recSHIP2 protein. We used Western blotting to determine the amounts of endogenous Ship2 protein in wild-type RCS cells (means ± SEM, 287,546 ± 37,305 molecules per cell; Fig. 6B). Indi-vidual cells from two RCS Ship2 null cell lines (Ship2c−/− and Ship2 g−/−) were microinjected with an amount of purified recSHIP2 equal to about 1/10 of the endogenous amount of Ship2 (~25,000 mol-ecules per cell), together with a dTomato transcriptional reporter for FGF-mediated ERK activation [pKrox24(MapErk)dTomato] (28). Cells were treated with FGF2, and the induction of dTomato expression was analyzed 24 hours later and plotted. Injection of recSHIP2 in-creased the ability of FGF to activate Erk in the Ship2Crispr cells (Fig. 6, C and D), demonstrating that addition of SHIP2 protein signifi-cantly increased the Erk activation by FGF signaling.

To test whether other RTKs also require Ship2 to activate Erk, we transfected RCS cells with constructs encoding full-length wild-type human tropomyosin receptor kinase A (TRKA) or human epi-dermal growth factor receptor (EGFR) and treated the cells with the TRKA ligand nerve growth factor (NGF) or the EGFR ligand EGF,

respectively. We then compared TRKA- and EGFR-mediated Erk activation be-tween wild-type cells and two Ship2Crispr cell lines. No substantial differences in Erk activation were found among wild-type and Ship2Crispr cells (fig. S5, A and B), suggesting that the requirement for Ship2 is specific to FGFRs.

FGFRs relay signaling to the ERK pathway through adaptors such as FRS2 and GAB1, both of which undergo tyro-sine phosphorylation upon FGF treat-ment and form complexes with the tyrosine phosphatase and adaptor pro-tein tyrosine–protein phosphatase non- receptor type 11 (PTPN11; also known as SHP2) to activate the RAS-ERK sig-naling module. FGF2 elicited abundant tyrosine phosphorylation of Frs2 at Tyr436 and Gab1 at Tyr627 in RCS cells (27), and this was impaired in Ship2Crispr cells (Fig. 6A and figs. S3, A to C, and S4, A to D). These data suggest that the defect in Erk activation in Ship2Crispr cells may stem from poor recruitment of Ptpn11 to underphosphorylated adap-tors. Despite multiple attempts, we were unable to immunoprecipitate Gab1 from RCS cells due to technical difficulties. However, Frs2- Ptpn11 coimmunopre-cipitation experiments showed impaired association of Frs2 with Ptpn11 upon FGF2 treatment of Ship2Crispr cells com-pared to wild-type cells (Fig. 6E and fig. S3D).

SHIP2 promotes association of SRC family kinases with FGFR3The SFKs BLK, FGR, FYN, HCK, LCK,

LYN, and YES were among the most frequently identified pro-teins in mass spectrometry analyses of the FGFR3 interactome (19), and coimmunoprecipitation experiments carried out in 293T cells confirmed the association of FGFR3 with LCK (Fig.  7A), FYN, LYN (fig. S6A), FGR, and BLK (fig. S6B). We used bimolecular fluorescence complementation (BiFC) to visualize the interaction of FGFR3 and LYN in RCS cells. For this analysis, LYN comple-mentary DNA (cDNA) was C-terminally tagged with cDNA encoding the N- terminal 158 amino acid residues of the yellow fluorescent protein (YFP) variant Venus (LYN-V1) (29), and FGFR3 cDNA was fused to cDNA encoding amino acid residues 159 to 239 from the C terminus of Venus (FGFR3-V2). When brought into close proximity, the N- and C-terminal Venus frag-ments reconstitute fluorescence. BiFC assays demonstrated that FGFR3 and LYN were located in very close proximity to one an-other in RCS cells (Fig. 7, B and C).

Inhibition of endogenous SFK catalytic activity by two unrelated chemical inhibitors (AZM475271 and A419259) abolished ERK ac-tivation in response to the addition of exogenous FGF2 in both RCS and 293T cells (Fig. 7, D and E). ERK inhibition was accompanied

A

100

– SHIP2-WT –

FGFR3-FLAG + – + +

+ – +

– – –

– + –

+ – –

SHIP2

WB

Input FGFR3

Actin

SHIP2

FGFR3 IP FLAG

SHIP2-∆PS

130

100

130

B kDa

C –

SHIP2-WT – FGFR3-FLAG +

– + +

+ +

– –

– + – + –

SHIP2-∆PR1 – – –

SHIP2

Input FGFR3

Actin

130

100

WB

FGFR3 IP FLAG

130

100

SHIP2

D –

SHIP2-N1 – FGFR3-FLAG +

– + +

+ – +

– – –

– + –

+ – –

Input

SHIP2-N2 kDa

SHIP2

FGFR3

55

35

WB

FGFR3 IP FLAG

35

IgH 55

Actin

kDa

SHIP2

E –

SHIP2-C1 – FGFR3-FLAG +

– + +

+ –

+ – –

– – + –

+ –

– kDa

SHIP2

FGFR3

Actin

70

55

35

WB

FGFR3

IgH

Input

SHIP2-C2

70

55

35

SHIP2

IP FLAG

F

– FGFR3-K508M-FLAG –

SHIP2 + – + +

+ – +

– – –

– + –

+ – –

Input

FGFR3-WT-FLAG

SHIP2

FGFR3

WB FGFR3

IP FLAG

SHIP2

Actin

Amino acids

Fig. 2. SHIP2 interacts with FGFR3 through its SH2 and SAM domains. (A) Schematic representation of truncated SHIP2 constructs. The V5 epitope was added to the C terminus of each construct. (B to E) The indicated SHIP2 variants were expressed with WT FGFR3 in 293T cells, immunoprecipitated (IP) using the Flag epitope tag, and analyzed by Western blotting (WB) for FGFR3 and SHIP2. Actin is a loading control. Input, total cell lysates used for IP. Data are representative of three independent experiments. IgH, immunoglobulin heavy chain. (F) Wild-type SHIP2 (SHIP2-WT)was coexpressed in 293T cells with WT FGFR3 (FGFR3-WT) or the catalytically inactive mutant FGFR3-K508M as indi-cated. Data are representative of three independent experiments.

on February 21, 2020

http://stke.sciencemag.org/

Dow

nloaded from

Page 5: CELL BIOLOGY Copyright © 2018 The inositol phosphatase ... · Fafilek et al., Sci. Signal. 11, eaap8608 2018 18 September 2018 SCIENCE SIGNALING| RESEARCH ARTICLE 1 of 14 CELL BIOLOGY

Fafilek et al., Sci. Signal. 11, eaap8608 (2018) 18 September 2018

S C I E N C E S I G N A L I N G | R E S E A R C H A R T I C L E

5 of 14

by the inhibition of FRS2 phosphorylation, suggesting that SFKs are recruited to FGFRs and participate in FGFR-mediated adaptor phosphorylation and ERK activation.

Having established that Ship2 deletion and SFK inhibition both impaired FGF-mediated ERK activation in human and rat cells, we speculated that SHIP2 promotes the association of SFKs with FGFRs. We used PLAs to probe the interaction of FGFR3 with the SFK YES in different Ship2Crispr RCS cell lines. Expression of FGFR3 in RCS cells caused it to become spontaneously activated (Fig. 7F) and associate with coexpressed YES at the cell membrane (Fig. 7, G and H). This association was reduced by 40 and 61% in two tested Ship2Crispr cell lines (Fig. 7I), suggesting that Ship2 facilitates the interaction of FGFR3 with SFKs. Next, we used the CRISPR/Cas9 system to disrupt the SHIP2 gene in U2OS cells expressing FGFR1-GFP and used PLA to test the association of FGFR1 with endogenous LYN. In the two SHIP2Crispr U2OS clones, the inter-action of FGFR1 with LYN was inhibited by 47 and 66%, compared to wild-type cells (Fig. 7, J to L). Together, the PLA experiments demonstrated less association of SFKs with FGFRs in cells with deleted SHIP2.

The inositol phosphatase activity of SHIP2 is not necessary for FGF-mediated ERK activationThe truncated human SHIP2 variant lacking the entire inositol phosphatase domain (SHIP2-PS) interacted nor-mally with FGFR3 (Fig. 2B). Similarly, FRS2, LYN, LCK, and FGR coimmu-noprecipitated normally with SHIP2-PS or with the phosphatase- defective SHIP2 (SHIP2-PD) triple mutant (P686A/ D690A/R691A; Fig. 8, A to C, and fig. S7) (30). Thus, the interaction with FGFR3 or with members of the FGFR3 signal-ing complex does not require SHIP2’s inositol phosphatase domain or its cat-alytic activity.

To test whether the catalytic func-tion is required for the role of SHIP2 in FGF-ERK signaling, RCS cells were treated with a chemical inhibitor of SHIP2 ino-sitol phosphatase activity (AS1949490) (31). Because the ability of SHIP2 to negatively regulate the phosphoinositide 3-kinase (PI3K)–AKT pathway depends on its catalytic activity, AS1949490 should increase AKT signaling. AS1949490 treatment produced no changes in basal phosphorylation of Akt at Ser473 in FGF2-naïve cells. However, AS1949490 prevented the inhibition of Akt activity by prolonged Fgfr signaling in RCS cells (Fig. 8D and fig. S8) (32), confirming that the inhibition of Ship2’s catalytic activity increased Akt activation. AS1949490 did not alter FGF-mediated FRS2 phosphorylation and, unexpect-edly, enhanced rather than suppressed FGF-mediated ERK phosphorylation

(Fig. 8D and fig. S8). Next, we monitored Erk activity in RCS cells using the pKrox24(MapErk)Luciferase reporter, which uses the promot-er of a locus strongly induced by ERK signaling (EGR1) to drive expression of a luciferase reporter. pKrox24 was designed to record quantitative changes in ERK pathway activation by FGF signaling (28). FGF-mediated pKrox24 transactivation was enhanced in cells transfected with wild-type human SHIP2, compared to untrans-fected controls (~13-fold versus ~9-fold; Fig. 8E). The pKrox24 transactivation was enhanced up to ~17-fold in cells transfected with SHIP2-PS, together suggesting that lack of SHIP2 catalytic activity does not account for diminished FRS2 phosphorylation and Erk activation in FGF-treated Ship2Crispr cells. Thus, the decrease in FGF-mediated Erk activation caused by loss of Ship2 does not result from the loss of Ship2 catalytic activity but likely reflects the role of Ship2 as an adaptor in the signaling cascade.

DISCUSSIONMost of the 20 established RTK families use the ERK pathway as a major effector of their signal transduction. The magnitude and

B

FGFR

1-G

FPFG

FR4-

GFP

CAGFPPLA Merged/DAPI

FGFR

1-G

FPFG

FR1-

BirA

-HA

131 93 101 77

***n = 3

***

Cells

PLA

eve

nts/

cell

GFPPLA Merged/DAPID

FGFR1-GFP

U2OS

FGFR1-BirA-HAFGFR4-GFP

GFP:SHIP2 FGFR1:SHIP2

++

+ +– –

– ––

FGFR-V5

SHIP2-FLAG

IP FLAG

– –

– – – –

Fig. 3. SHIP2 interacts with FGFR1 in cells. (A) 293T cells were cotransfected with FLAG-tagged SHIP2 and V5-tagged WT FGFR1, FGFR2, or FGFR4, and lysates were subjected to FLAG IP and Western blotting (WB) for the FGFRs and SHIP2. Data are representative of three independent experiments. (B to D) U2OS cells stably expressing FGFR1-GFP were subjected to PLA using antibodies detecting SHIP2 and GFP (B and C) or SHIP2 and FGFR1 (B and D). U2OS cells stably expressing FGFR1-BirA-HA or FGFR4-GFP were used as negative controls in (B) and (C) and (B) and (D), respec-tively. n = 3 independent experiments. Data are means ± SEM. Statistical significance was determined by Welch’s t test with Bonferroni’s correction of P values (n.s., P > 0.05; ***P < 0.001). Scale bar, 20 m. Numbers in columns (B) indicate the total number of cells scored across the three independent experiments.

on February 21, 2020

http://stke.sciencemag.org/

Dow

nloaded from

Page 6: CELL BIOLOGY Copyright © 2018 The inositol phosphatase ... · Fafilek et al., Sci. Signal. 11, eaap8608 2018 18 September 2018 SCIENCE SIGNALING| RESEARCH ARTICLE 1 of 14 CELL BIOLOGY

Fafilek et al., Sci. Signal. 11, eaap8608 (2018) 18 September 2018

S C I E N C E S I G N A L I N G | R E S E A R C H A R T I C L E

6 of 14

duration of ERK activation appear critical for cell decisions during the physiological processes regulated by RTKs. In PC12 pheochromocytoma cells, for instance, the transient Erk activa-tion elicited by Egfr leads to increased proliferation, whereas the sustained Erk activation elicited by Trk results in growth arrest and neuronal differentiation (33, 34). Under pathological condi-tions, sustained ERK activation is almost invariantly the major driver of cell phenotypes caused by RTK deregulation. The

FGFRs are no exception from this para-digm, but the mechanisms associated with sustained ERK activation in FGFR signaling have been unclear. Deepening our understanding of the mechanisms associated with ERK activation down-stream of FGFRs will aid in progress toward treatments for FGFR-related pathologies.

FGFRs relay their signal to ERK through adaptors such as FRS2. FGFRs phosphorylate FRS2 at several tyrosine residues that serve as binding sites for the GRB2-SOS1 (Tyr196, Tyr306, Tyr349, and Tyr392) and PTPN11-GRB2-SOS1 (Tyr436 and Tyr471) complexes, thus directing SOS1 guanine nucleotide ex-change factor activity to the cell mem-brane where it activates RAS (16, 17). Adaptors not only relay the signal but also act as its amplifiers because the phosphorylation of several adaptor molecules by a single FGFR substan-tially increases the docking interface for cytoplasmic SOS1 at the cell mem-brane. Thus, RTK families that use adaptors, such as FGFR or TRK, tend to produce higher and longer ERK acti-vation compared to RTKs that engage their downstream signaling targets directly (35).

Although FGFRs induce strong ERK activation in general, the magnitude and duration of ERK activity fluctuate sub-stantially among different cell types, ranging between 1 and 2 hours in fibro-blasts to persistent signal for at least 24 hours in chondrocytes (36). The mechanisms maintaining this ERK activity are poorly understood. Accord-ing to the existing paradigm, activation of FGFRs leads to their rapid internal-ization and degradation. This is followed by a period during which a given cell is insensitive to FGF, before new FGFR molecules are present at the cell surface (4, 37–39). In some pathological con-texts, the continuous presence of FGFRs at the cell surface contributes to sus-tained ERK activity. These include gas-tric and lung cancers, in which FGFR1

and FGFR2 are frequently amplified, leading to constant gene ex-pression and protein production, and multiple myeloma, in which FGFR3 is overproduced as a result of the t(4;14)(p16.3;q32) chromosomal translocation that places an immunoglobulin pro-moter upstream of the FGFR3 gene (40–42). Although the high abundance of FGFR3 is part of the chondrocyte differentiation program, experimental evidence has demonstrated rapid FGFR3 down-regulation after its activation even in these cells (25). Thus,

FGF2 (hours) – 1 2

Ship2

A

D

FGF2

Control pp130Cas pp130Cas/vinc./DAPI

pp13

0Cas

/vin

culin

pS

hip2

/vin

culin

E

pShip2 pShip2/DAPI pShip2/DAPI/DIC

FGF2

Control B

pShip2 pShip2/vinc./DAPI Control

C

FGF2

F G

pShip2Y986/7

Fig. 4. FGF signaling targets SHIP2 to focal adhesions. (A) RCS cells were treated with FGF2 for the indicated times, and lysates were subjected to immunoblotting for Ship2 phosphorylated on Tyr986 or Tyr987 (pShip2Y986/7) and for total Ship2. Data are representative of three independent experiments. (B) Immunofluorescence images showing phosphorylated Ship2 (pShip2Y986/7) in RCS cells treated with FGF2 or not (control). Arrows indicate peripheral focal adhesions. DAPI, 4′,6-diamidino-2-phenylindole; DIC, differential interference contrast. (C) Immunofluorescence im-ages showing pShip2Y986/7 and the focal adhesion marker vinculin (vinc.). Arrows indicate focal adhesions in RCS cells treated with FGF2. (D) Immunofluorescence images showing p130Cas phosphorylated at Tyr410 (pp130Cas) and vinculin in RCS cells treated with FGF2. Arrows indicate focal adhesions. Scale bars, 25 m. (E) Higher magnification of boxed areas indicated in (C) and (D). Images are representative of three independent experiments. (F) Immunoblot showing Ship2 in the indicated Ship2Crispr RCS cell lines compared to WT RCS cells. Data are representative of three independent experiments. (G) Quantification of fetal bovine serum (FBS)– and FGF2-stimulated migration in WT RCS cells and the indicated Ship2Crispr cell lines. Data are means ± SEM. Statistical significance was determined by Welch’s t test with Bonferroni’s correction of P values, ***P < 0.0001. Data are representative of three independent experiments.

on February 21, 2020

http://stke.sciencemag.org/

Dow

nloaded from

Page 7: CELL BIOLOGY Copyright © 2018 The inositol phosphatase ... · Fafilek et al., Sci. Signal. 11, eaap8608 2018 18 September 2018 SCIENCE SIGNALING| RESEARCH ARTICLE 1 of 14 CELL BIOLOGY

Fafilek et al., Sci. Signal. 11, eaap8608 (2018) 18 September 2018

S C I E N C E S I G N A L I N G | R E S E A R C H A R T I C L E

7 of 14

the persistence of FGFRs at the membrane is unlikely to account for sustained ERK activation in chondrocytes.

The adaptors that mediate ERK activation downstream of FGFRs may attenuate the ERK signal. FRS2 represents such a site of potent negative feedback. First, tyrosine-phosphorylated FRS2 mediates the interaction of the ubiquitin ligase cCBL with FGFR, leading to its ubiquitin-mediated degradation (43). Second, active ERK phosphorylates FRS2 at multiple threonine residues adjacent to the phosphorylated tyrosines, leading to diminished PTPN11-GRB2-SOS1 recruitment and attenuation of ERK activity (44, 45).

The components of the ERK-FRS2 negative feedback loop exist in chon-drocytes, yet there is no corresponding attenuation of ERK activity in these cells (27). Moreover, analyses of transcrip-tional changes triggered by FGFR acti-vation in chondrocytes (18) demonstrated a potent induction of the ERK phos-phatase dual specificity phosphatase 6 and members of the Sprouty family of RAS-ERK pathway inhibitors (46, 47). Thus, FGF-mediated ERK activation persists for many hours despite the induction of negative feedback mecha-nisms, suggesting the existence of other molecular processes capable of over-riding established negative feedback mechanisms. Our study describes one such mechanistic pathway, linking the PS SHIP2 to the FGFR signaling complex for sustained ERK activation. We showed that SHIP2 interacted with FGFRs, was phosphorylated by active FGFRs, and promoted FGFR-ERK sig-naling at the level of adaptor phos-phorylation and PTPN11 recruitment (Fig. 8F).

The SFKs are known to participate in a wide array of FGF-regulated events including cell proliferation, shape changes, migration, adhesion, and differentiation (48–55). The mechanism of FGFR- mediated recruitment of SFKs is some-what unclear, although earlier studies demonstrated direct interaction of SFKs with FGFRs, or recruitment of SFKs by FRS2 (56–58). Similarly, although most of the cellular phenotypes caused by FGF-SRC signaling depend on ERK (48, 59), it is not clear how SFKs mech-anistically integrate into the FGF-ERK pathway. We showed that SFKs are re-cruited to the FGFR signaling complex by SHIP2. Inhibition of SFK activity abolished FGF-mediated ERK activa-tion and FRS2 phosphorylation, suggest-ing that SFKs assist FGFRs in adaptor phosphorylation and signal relay to the ERK pathway, together enabling a ro-

bust, negative feedback–resistant relay of the FGFR signal to the RAS-ERK pathway. Our experiments thus establish SHIP2 as an essential component of canonical FGFR signaling that enables sustained FGFR-mediated ERK activation by recruiting SFKs to the FGFR signaling complex (Fig. 8F).

The major physiological function attributed to SHIP2 is the regulation of insulin signaling. Ship2 knockout mice show increased sensitivity to insulin, causing hypoglycemia, deregulated expres-sion of genes involved in liver gluconeogenesis, and perinatal death (60). A later study did not confirm such a severe phenotype

A

C

E F G

D

B

Fig. 5. Loss of SHIP2 rescues some FGF-induced cell phenotypes. (A) Quantification of cell proliferation, as deter-mined by crystal violet staining, in WT RCS cells and the four indicated Ship2Crispr RCS cell lines in the presence of in-creasing amounts of FGF2. (B) Quantification of cell proliferation in three randomly selected RCS clones in which the Ship2 was not disrupted by CRISPR/Cas9 (2E11, 1E11, and 2B10) and the three indicated Ship2Crispr RCS cell lines. Data in (A) and (B) represent averages from eight wells ± SD. Results are representative of three independent experiments. Statistically significant differences are highlighted. ***P < 0.001, Welch’s t test. (C) Immunoblot showing total Ship2 in the indicated RCS cell lines. Results are representative of three independent experiments. (D) Immunoblot show-ing the cartilaginous ECM marker collagen 2 and the senescence marker caveolin in two WT and two Ship2 knockout RCS cell lines in the presence and absence of FGF2. Blot is representative of three independent experiments. (E) Alcian blue staining showing sulfated proteoglycans characteristic of cartilaginous ECM in the indicated WT and Ship2 knockout RCS cell lines in the absence and presence of FGF2. Scale bar, 200 m. (F) Evidence for spreading in cells treated with FGF2 for 72 hours, due to the formation of actin stress fibers, visualized by phalloidin staining. Scale bar, 200 m. (G) Immunofluorescence showing phalloidin in WT and Ship2Crispr cells in the absence and presence of FGF2. Scale bars, 20 m. Data are representative for three independent experiments.

on February 21, 2020

http://stke.sciencemag.org/

Dow

nloaded from

Page 8: CELL BIOLOGY Copyright © 2018 The inositol phosphatase ... · Fafilek et al., Sci. Signal. 11, eaap8608 2018 18 September 2018 SCIENCE SIGNALING| RESEARCH ARTICLE 1 of 14 CELL BIOLOGY

Fafilek et al., Sci. Signal. 11, eaap8608 (2018) 18 September 2018

S C I E N C E S I G N A L I N G | R E S E A R C H A R T I C L E

8 of 14

but reported that Ship2−/− animals were resistant to weight gain due to the decreased cellular response to insulin (61). In addition to altered insulin signaling, the Ship2−/− mice are about half the size of their wild-type littermates, suggesting a potential role for SHIP2 in the growth of both the axial and appendicular skele-

tons. In humans, loss-of-function mu-tations in SHIP2 cause OPS, an auto-somal recessive skeletal dysplasia characterized by severe shortening of all the appendicular bones with radio-graphic evidence of endochondral ossification delay, extremely short hands, poor bone mineralization, flat-tening of the spine, and severe midface hypoplasia (62–64). Histologic analy-ses of cartilage growth plate morphol-ogy in patients with mutations in SHIP2 showed poor organization of the zone of proliferating chondrocytes, near the absence of the zone of hypertrophic chondrocytes, and increased vascular invasion (65, 66).

Molecular analyses have demonstrated that the negative role of SHIP2 in insulin signaling is caused by SHIP2- mediated inhibition of PI3K-AKT signaling (30, 61). SHIP2-mediated dephosphorylation of plasma membrane PtdIns(3,4,5)P3 to PtdIns(3,4)P2 reduces the amount of AKT that can be recruited to the cell mem-brane and subsequently activated by PDK1 (3-phosphoinositide- dependent protein kinase-1) (67). Similar to the role of SHIP2 in insulin signaling, the develop-mental defects in OPS result from the lack of SHIP2 catalytic activity because several missense mutations associated with OPS map to the catalytic domain (68, 69). Because insulin-like growth factor (IGF) signals primarily through the PI3K-AKT pathway (70), it is likely that the mutations in SHIP2 that cause OPS do so through deregulation of IGF- AKT mitogenic signaling. IGF signaling is a major growth-promoting signal for skeleton (71). Knockout models of IGF or its receptor IGF-1R show growth re-tardation with poor bone mineraliza-tion and an abnormal cartilage growth plate that shares phenotypic similari-ties with the human OPS cases charac-terized by disorganized proliferating chondrocytes and a shortened hyper-trophic chondrocyte zone, supporting that the hypothesis abnormal IGF sig-naling affects the growth plate (72, 73). In addition to PI3K-AKT signaling, IGF-1R interacts with the parathyroid hormone–related peptide–Indian hedge-

hog pathway, adding additional complexity to direct and indirect effects of SHIP2 loss on the cellular behavior of chondrocytes that are also affected by dysregulation of FGFR signaling (73).

However, the SHIP2-mediated propagation of FGF signaling in chondrocytes appears to be independent of SHIP2 catalytic activity

A 8 1042–

WT61 842–

Ship2

a+/–

61

Gab1

FGF2 (hours)

pErkT202/Y204

Erk

Akt

pAktS473

pGab1Y627

pFrs2Y436

Frs2

10D

IP Frs2 Frs2

Ptpn11

Ptpn11

Frs2

WB

Actin

6 – WT 4

Ship2

a+/–

FGF2 (hours) 6 – 4 Input

B

C

E

Actin

Ship2g–/–

Ship2c–/–

Fig. 6. Loss of Ship2 impairs Fgfr-Erk signaling. (A) Representative Western blot showing total and phosphorylated (p) forms of Frs2, Gab1, Akt, and Erk in WT and Ship2Crispr (Ship2a+/−) RCS cells treated with FGF2 for the indicated times. Actin is a loading control. Densitometry quantification and number of independent replicates are provided in fig. S4A. Additional analyses carried out in other Ship2Crispr cell lines are shown in figs. S3 (A to C) and S4 (C and D). (B) Quantification of Ship2 protein in three WT RCS cultures by Western blotting of lysates and serial dilutions of purified recombinant (rec) SHIP2, followed by densitometry to quantify the band intensities. The number of Ship2 molecules per cell was estimated on the basis of comparing the band intensities and number of cells in each lysate (lanes

1 to 3) with the band intensities of the purified protein samples (lanes 4 to 10). The numbers above lanes 1 to 3 indi-cate the number of cells represented by the lysates, and the numbers above lanes 4 to 10 represent the number of molecules of recSHIP2 per sample. On the basis of these calculations, we estimated there to be 287,546 ± 37,305 (means ± SEM) Ship2 molecules per cell. n = 9 independent RCS cultures. I.O.D., integrated optical density. (C) Ship2Crispr cells were microinjected with amounts of recombinant SHIP2 (recSHIP2) equal to about 1/10 (~25,000 molecules per cell) of the number of endogenous Ship2 molecules per cell along with a dTomato transcriptional re-porter of ERK activity [pKrox24(MapErk)dTomato]. TexasRed conjugated to dextran was used as a marker for injection. The cells were then treated with FGF2 and assayed for dTomato expression. Ph2, phase contrast. Scale bar, 200 m. (D) Quantification of dTomato expression in the two indicated Ship2Crispr cell lines treated as in (C). n = 3 indepen-dent experiments. ***P ˂ 0.001, **P ˂ 0.01, Student’s t test. (E) WT and Ship2a+/− RCS cells were treated with FGF2 for the indicated times, Frs2 was immunoprecipitated (IP) from the lysates, and the immunoprecipitates were sub-jected to Western blotting for Ptpn11. Blot is representative of three independent experiments.

on February 21, 2020

http://stke.sciencemag.org/

Dow

nloaded from

Page 9: CELL BIOLOGY Copyright © 2018 The inositol phosphatase ... · Fafilek et al., Sci. Signal. 11, eaap8608 2018 18 September 2018 SCIENCE SIGNALING| RESEARCH ARTICLE 1 of 14 CELL BIOLOGY

Fafilek et al., Sci. Signal. 11, eaap8608 (2018) 18 September 2018

S C I E N C E S I G N A L I N G | R E S E A R C H A R T I C L E

9 of 14

and the AKT pathway. First, deletion or reduction of Ship2 in Ship2Crispr RCS cells did not increase AKT signaling as would be expected whether the role of Ship2 is limited to the inhibition of the PI3K-AKT pathway. Second, chemical inhibition of Ship2 phosphatase activity increased AKT phosphorylation but did not reduce FGF-mediated ERK activation or adaptor phosphorylation. Third, genetic inactivation or deletion of the SHIP2 catalytic domain did not affect its association with the FGFR signaling

complex. It thus appears that SHIP2’s catalytic function is im-portant for mitogenic IGF signaling in growth plate cartilage, whereas its adaptor function is critical for the growth-inhibitory signaling of FGFR3 in the same tissue. Further experiments should unravel the interesting yet unknown biology behind

+ +

+ – – + FGFR3

AZD (100 nM)

pFGFR3Y653/4

FGFR3

Actin

– –

DAPI/YES-YFP/PLA

z y

ERK

– – +

0.5 + 1

+ 0.1

– 5

– – +

– + –

– –

pERKT202/Y204

5 10 20 20 – + + + – + +

– – –

FGF2 SRC inh. (µM)

– –

FRS2

pFRS2Y436

D

A

E

YES-YFP

WT

PLA H

F

G

Sh

ip2

a+

/–

S

hip

2

c–/– YFP only

FGFR3 YES-YFP

Empty vector YFP only

Empty vector

Ship2

a+/–

(n = 3) WT (n = 4) Ship2

c–/–

(n = 3) I

YES-YFP FGFR3 Transfection

YES-YFP

DAPI

FGFR

3-Y

ES

inte

ract

ion

(x10

–6)

(nor

mal

ized

to Y

ES

/YFP

exp

ress

ion)

LYN-V1 FGFR3-V2

Ven

us-p

ositi

ve c

ells

(%)

Cells counted

n = 3

+ + – + + –

122 120 91

C B

Ven

us fl

uor.

Ven

us IC

C

Mer

ged/

DA

PI

FGFR3-V2 LYN-V1 FGFR3-V2 LYN-V1

AZM475271 A419259

K J

SHIP2b–/–

WT

FGFR4-GFP

SHIP2g–/–

SH

IP

2b–/–

FGFR

4-G

FP

U2OS

+

+ + –

+ – – – –

– – – –

– – –

RC

S

SH

IP

2g–/–

WT

SH

IP

2

g–/–

SH

IP

2

b–/–

1RF

GF:S

O2U

-GFP

U

2OS:

FG

FR4-

GF P

L

Cells 258 183 294 310

n = 3 *** **

Cells

FGFR

1-G

FP U2OS

FGFR

1-G

FP

FGFR3-WT-FLAG FGFR3-K650E-FLAG

IP FLAG

LCK

Actin

FGFR3

LCK Input

FGFR3

LCK WB

M

Fig. 7. SHIP2 promotes the association of SFKs with FGFR3. (A) 293T cells were transfected with vectors encoding FLAG-tagged WT FGFR3 or the K650E activating mutant (FGFR3-K650E) together with the SFK LCK. FGFR3 was immuno-precipitated (IP) from lysates using the FLAG tag, and the immunoprecipitates were analyzed for FGFR3 and LCK by Western blotting (WB). (B and C) Bimolecular complementation assay. (B) RCS cells were transfected with vectors encoding FGFR3-Venus2 (V2) and Lyn-Venus1 (V1), fixed, and counterstained by GFP antibody to visualize expression of either part of Venus protein. ICC, immunocytochemistry; fluor., fluorescence. (C) The percentage of transfected cells showing Venus activity was calculated and plotted. n = 3 independent experiments. Data represent aver-ages ± SEM. Scale bar, 10 m. (D and E) RCS cells (D) and 293T cells (E) were treated with FGF2 alone or in the presence of the SRC inhibitors (inh.) AZM475271 or A419259 and analyzed by Western blotting for total and phosphorylated (p) FRS2 and ERK. Data are representative of three independent experiments. (F) Western blot show-ing activation (phosphorylation) of WT FGFR3 overexpressed in RCS cells in the absence and presence of the inhibitor of FGFR3 catalytic activity AZD4547 (AZD). (G and H) Immunofluo-rescence images showing PLAs for FGFR3 and the SFK YES tagged with YFP (YES-YFP) in RCS cells. Scale bars, 10 m. (I) Combined results of three independent PLA experiments assessing the co-localization of FGFR3 and YES in two Ship2Crispr cell lines (Ship2c−/− and Ship2a+/−) compared to WT RCS cells. Transfections containing empty vector and a vector expressing YFP only are negative controls for the PLA assay. Data represent aver-ages ± SEM. *P < 0.05, ***P < 0.001, Welch’s t test with Bonferroni’s correction of P values. (J) Immu-nofluorescence images showing PLAs testing the association of plasmid-encoded FGFR1-GFP with the endogenous SFK LYN. (K) Western blot for SHIP2 in U2OS clones with the indicated SHIP2 genotypes. (L) Combined results of three inde-pendent PLA experiments assessing the colocal-ization of FGFR1 and LYN in two SHIP2Crispr U2OS cell lines (SHIP2g−/− and SHIP2b−/−) compared to SHIP2+/+ U2OS cells. U2OS cells expressing FGFR4-GFP were used as a negative control for the PLA assay. Data represent averages ± SEM (**P < 0.01, ***P < 0.001, Welch’s t test).

on February 21, 2020

http://stke.sciencemag.org/

Dow

nloaded from

Page 10: CELL BIOLOGY Copyright © 2018 The inositol phosphatase ... · Fafilek et al., Sci. Signal. 11, eaap8608 2018 18 September 2018 SCIENCE SIGNALING| RESEARCH ARTICLE 1 of 14 CELL BIOLOGY

Fafilek et al., Sci. Signal. 11, eaap8608 (2018) 18 September 2018

S C I E N C E S I G N A L I N G | R E S E A R C H A R T I C L E

10 of 14

how these two growth factor systems, IGF and FGF, integrate SHIP2 into their downstream signaling to achieve their opposing effects on cell proliferation. These studies should also illuminate the therapeutic potential of modulating SHIP2 activity in devel-opmental pathologies caused by activating mutations in FGFRs.

MATERIALS AND METHODSCell culture, vectors, transfection, and reagents293T (American Type Culture Collection), U2OS, and RCS cells (obtained from B. de Crombrugghe) (74) were propagated in Dulbecco’s

modified Eagle’s medium (DMEM), supplemented with 10% fetal bo-vine serum and antibiotics (Invitrogen). All expression vectors are listed in table S1. To obtain cells stably expressing FGFR4-GFP, U2OS cells were transfected with pEGFP- N1-FGFR4 (75) using FuGENE6 according to the manufacturer’s protocol (Promega). Clones were selected with geneticin (1 mg/ml), and their FGFR4- GFP expression amounts were analyzed by immunofluorescence and Western blot. A homogenous clone with moderate FGFR4-GFP expression was chosen for further studies. U2OS cells stably expressing FGFR1-GFP and FGFR1-BirA-HA have been described (76). For generation of truncated SHIP2 variants, the C-terminal

A C

AS19 (10 µM) FGF2 (hours) 1

+ +2 4 6 1 2 4 6

+ + +8 10 8 10

+ +pFrs2Y436

Frs2

pErkT202/Y204

Erk

pAktS473

Akt

Actin

D

+ ++LCKInput

SHIP2 WT V5SHIP2 PD V5

+ +++

LYN

Actin

SHIP2

InputLYN

SHIP2WB

LYN

SHIP2 WT V5

+++

++

++

IP V5

B

GFP SHIP2 ∆PS V5

130

100

E

FGF2

med

iate

d in

duct

ion

(fold

cha

nge)

FRS2

SHIP2 WT V5

+++

++

++

kDaSHIP2 ∆PS V5

FRS2Input

130

100

kDa

130

100 WB

FRS2IP FRS2

130

100

SHIP2

SHIP2

Actin

F

LCK

Actin

SHIP2

WBSHIP2

IP V5

LCK

Fig. 8. The inositol phosphatase activity of SHIP2 is not necessary for its association with FGFR signaling complexes or FGF-mediated ERK activation. (A to C) Representative Western blots showing LCK (A), LYN (B), and FRS2 (C) in SHIP2-V5 coimmunoprecipitates (IP) from 293T cells expressing the indicated combinations of V5-tagged WT SHIP2, the catalytically inactive SHIP2 triple mutant SHIP2-PD (P686A, D690A, and R691A), a SHIP2 mutant lacking the entire inositol phosphatase domain (SHIP2-PS), LCK, LYN, and FRS2. Nontransfected cells or those transfected with vectors encoding only GFP serve as negative IP controls. Actin is a load-ing control. (D) RCS cells were treated with inhibitor of SHIP2 phosphatase activity AS1949490 (AS19) and purified FGF2 for the in-dicated times, and the lysates were analyzed by Western blotting for phosphorylated (p) forms of the indicated proteins. Actin and total protein are loading controls. Immuoblotting data are rep-resentative of three independent experiments. (E) Dual-luciferase assay. RCS cells were trans-fected with WT SHIP2 or SHIP2-PS together with a firefly luciferase pKrox24 ERK reporter plasmid that is transactivated by FGF-ERK signaling and a Renilla luciferase pTK-RL control plasmid. Cells were treated with FGF2 for 24 hours before the dual-luciferase assay. Data are compiled from three independent experiments, with four bio-logical and two technical replicates for each treatment. Bars are averages ± SD. **P < 0.01, ***P < 0.001, Welch’s t test with Bonferroni’s correction of P values. (F) Model illustrating the putative role of SHIP2 in canonical FGFR-ERK signaling. Upon being activated by binding to FGFs, FGFRs are activated and phosphorylate the adaptor protein FRS2. Although GAB1 phos-phorylation can be also detected in FGF-treated cells, FRS2 is the major adaptor involved in signal relay from FGFR to the ERK pathway (17, 45, 81). Activated FRS2 and GAB1 recruit PTPN11 and SOS1 to the membrane. This activates RAS, which in turn activates ERK. Activated FGFRs also phos-phorylate SHIP2 and recruit SHIP2 to the FGFR signaling complex at the cell membrane. SHIP2 recruits SFKs to the complex, which enhances FGFR-mediated phosphorylation of the adaptor proteins, activation of RAS, and ERK signaling. SHIP2 knockout effectively converts FGF-induced sustained ERK activation into the FGF-induced transient activation and rescues the cell phenotypes induced by sustained FGFR-ERK signaling (premature senescence, ECM degradation, and growth arrest). This is due to hypophosphorylation of FRS2 in SHIP2 knockout cells, resulting in diminished recruitment of PTPN11-SOS1 complexes that activate the RAS-ERK signaling module.

on February 21, 2020

http://stke.sciencemag.org/

Dow

nloaded from

Page 11: CELL BIOLOGY Copyright © 2018 The inositol phosphatase ... · Fafilek et al., Sci. Signal. 11, eaap8608 2018 18 September 2018 SCIENCE SIGNALING| RESEARCH ARTICLE 1 of 14 CELL BIOLOGY

Fafilek et al., Sci. Signal. 11, eaap8608 (2018) 18 September 2018

S C I E N C E S I G N A L I N G | R E S E A R C H A R T I C L E

11 of 14

Myc-DDK tag in pCMV6-hSHIP2 vector (OriGene) was replaced by V5-HIS tag using a NEBuilder HiFi DNA assembly kit (New England Biolabs). Truncated human SHIP2 constructs were gener-ated by polymerase chain reaction (PCR) mutagenesis. Catalytically inactive 5′-PD SHIP2 was created by introducing P686A, D690A, and R691A substitutions into the inositol phosphatase domain by site-directed mutagenesis (Agilent). For cell migration assay, cells were plated on fibronectin-coated eight-chambered -slide wells and allowed to migrate for 15 hours as reported before (23, 77). Cells were analyzed with a Leica DM6000B microscope, using 10× magnification objective for live-cell imaging. Each cell was tracked over the period of time using the manual tracking plugin of the ImageJ software. FGF2, NGF, and EGF were from R&D Systems, heparin was from Sigma-Aldrich, AZM475271, A419259, and AS1949490 were from Tocris Bioscience.

Bimolecular fluorescence complementationFor the BiFC assays, cDNA encoding amino acid residues 2 to 158 of Venus was fused to the C terminus of LYN cDNA (LYN-V1) (29), and cDNA encoding amino acid residues 159 to 239 of Venus was fused to the C terminus of FGFR3 cDNA (FGFR3-V2). The ECM of RCS cells was degraded by 4-hour treatment with 0.3% bacterial collagenase 2 (Gibco) in no serum DMEM. Cells were transfected using FuGENE6 (1 g of DNA per well of 24-well plate), fixed using 4% paraformaldehyde 24 hours later, and coun-terstained for transfection using GFP antibody (table S2), and the per-centage of Venus complementation in transfected cells was calculated.

Dual-luciferase assaysFor dual-luciferase assay, the pKrox24(2xD-E_inD)Luc firefly lu-ciferase reporter (28) was transfected together with the pTK-RL vector (Promega) in 3:1 (microgram of DNA) ratio. Both plasmids were transfected using FuGENE6 reagent, 0.75 g of plasmid DNA per well of a 24-well plate containing 2 × 105 RCS cells, in 1:3.2 ratio of DNA/FuGENE6 (micrograms per microliter). The cells were treated with FGF2 for 24 hours, and the luciferase signals were de-termined using dual-luciferase assay (Promega). Briefly, cells were lysed in passive lysis buffer (100 l per well) for 15 min, and 10 l of lysate was used for determination of the dual-luciferase signal using the manufacturer’s protocol and TriStar Multimode Luminometer (Berthold Technologies); 40 l of each luciferase substrate was used per sample, and the measurement times were 5 to 10 s for firefly and 5 s for Renilla luciferase.

MicroinjectionCells were microinjected using a FemtoJet 4i microinjector with mi-cromanipulator InjectMan 4 (Eppendorf) with 80 hPa injection pres-sure, 0.3 s injection time, and 10 hPa compensation pressure. Each cell was injected with 200 fl of phosphate-buffered saline containing ~25,000 molecules of recSHIP2 (SignalChem) together with fluores-cence marker Dextran Alexa Fluor 647 (Thermo Fisher Scientific) and 100 molecules of pKrox24(MapErk)dTomato reporter. Cells were treated for 24 hours with FGF2 (10 ng/ml), fixed with 10% paraformal-dehyde, and immediately imaged for pKrox24 transactivation, using the automatic microscope TissueFAXS i (TissueGnostics) with 20× air ob-jective. We acquired images for red and far-red fluorescence channels showing pKrox24(MapErk)dTomato reporter and far-red counterstain marker Dextran Alexa Fluor 647. Percentages of injected cells positive for dTomato were determined by Fiji free software.

IP and Western blottingAbout 5 × 106 of transfected cells were lysed for 30 min in 1 ml of IP buffer containing 50 mM tris-HCl (pH 7.4), 150 mM NaCl, 0.5% NP-40, 0.25% sodium deoxycholate, 2 mM EDTA, and 1 mM Na3VO4, supplemented with protease inhibitors. For IPs, lysates were cleared by centrifugation (14,000g for 30 min), and supernatants were incu-bated for 1 hour with antibodies on rotating platform at 4°C. Im-munocomplexes were collected on 30 l of A/G agarose (Santa Cruz Biotechnology) in overnight incubation, spun down, and washed four times in 400 l of the IP buffer. Proteins attached to the beads were eluted to 70 l of loading buffer. For Western blotting, 20 l of cell lysates or immunoprecipitates were resolved by SDS-PAGE, transferred onto a polyvinylidene difluoride membrane, blocked with 5% (w/v) in tris-buffered saline (TBS)/Tween 20] nonfat milk, incubated with primary antibodies overnight at 4°C, washed 3× for 10 min in TBS/Tween 20, incubated with secondary antibodies at room temperature for 1 hour, washed 3× for 10 min TBS/Tween 20, and visualized by chemiluminiscence using Pierce ECL (Thermo Fisher Scientific), Immobilon Western (Millipore), or SuperSignal West Femto (Thermo Fisher Scientific) substrates. Table S2 lists all antibodies used in the study.

Kinase activity assaysTwo hundred nanograms of recombinant FGFR3 was incubated for 30 min at 30°C with recSHIP2 (SignalChem) as a substrate in 50 l of kinase buffer [60 mM Hepes (pH 7.5), 3 mM MgCl2, 3 mM MnCl2, 10 M Na3VO4, and 1.2 mM dithiothreitol] supplemented with 10 M ATP. For the SHIP2 activity assay, 300 ng of FGFR3 and 400 ng of SHIP2 were incubated with 50 M PtdIns(3,4,5)P3 (Echelon Biosciences) for 30 min at 30°C. The reaction was then incubated with BIOMOL Green Reagent (Enzo Life Sciences), and optical density at 620 nm was determined by a spectrophotometer. Crystal violet staining was carried out as described (78).

CRISPR/Cas9SHIP2 ablation in RCS and U2OS cells using CRISPR/Cas9 tech-nology was carried out as described (79). The CHOPCHOP tool was used to design single-guide RNAs for a pair of SpCas9n (D10A) nickases, which targeted 5′-GTGTGGGGCACCGAGTC-CCG-3′ and 5′-GTCACGGTGATACCAGGCAG-3′ sites in the first exon of rat Ship2 and 5′-CGATGGCAGCTTCCTGGTCC-3′ and 5′-GCGCTCTGCGTCCTGTGAGT-3′ sites in the first exons of the human SHIP2 gene (80). Two of each Crispr/Cas9n plas-mids (3 g; Addgene plasmid #42335) were electroporated (Neon Transfection System, Thermo Fisher Scientific; 100 l of tip, RCS cell protocol: 1600V, 10 ms, 3 pulses; U2OS cells: 1050 V, 30 ms, 2 pulses) together with a 0.5 g of GFP-based CRISPR Reporter [designed GFP-based reporter plasmid for CRISPR (GREP), a gift from V. Korinek] where the targeted sequence was cloned within the GFP coding sequence. Successful targeting restored the GFP-reading frame, and GFP- positive cells were manually picked and expanded. Individual clones were screened by West-ern blot for SHIP2, and the targeted locus was PCR-amplified using 5′-agcctccactccaagcttcc-3′ and 5′-aaggtctccactcacggt-gg-3′ primers for rat locus and 5′-GTACCACCGCGACCT-GAG-3′ (forward) and 5′-cttggctttctcctgggtct-3′ (reverse) primers for human locus. PCR products were cloned into pGEM-T easy vector (Promega) and sequenced to determine the SHIP2 genotype.

on February 21, 2020

http://stke.sciencemag.org/

Dow

nloaded from

Page 12: CELL BIOLOGY Copyright © 2018 The inositol phosphatase ... · Fafilek et al., Sci. Signal. 11, eaap8608 2018 18 September 2018 SCIENCE SIGNALING| RESEARCH ARTICLE 1 of 14 CELL BIOLOGY

Fafilek et al., Sci. Signal. 11, eaap8608 (2018) 18 September 2018

S C I E N C E S I G N A L I N G | R E S E A R C H A R T I C L E

12 of 14

Immunofluorescence and immunocytochemistryCells were fixed in 4% paraformaldehyde and incubated with pri-mary antibodies at 4°C overnight. Secondary antibodies were con-jugated with Alexa Fluor 488/594 (Life Technologies). Polymerized actin was visualized using Alexa Fluor 594–conjugated Phalloidin (Life Technologies) and vinculin-stained with vinculin–fluorescein isothiocyanate antibody according to the manufacturer’s protocol (Sigma-Aldrich). Images were taken using a confocal-inverted microscope Carl Zeiss LSM 700 with 63× oil immersion objective. Images shown in Figs. 3 (C and D), 4 (B to E), 5G,  and 7 (B, H, and J) represent maximum projections of several acquired z sections. Table S2 lists all antibodies used in the study.

Proximity ligation assayFor Duolink PLA (Sigma-Aldrich), cells were fixed with 4% para-formaldehyde for 15 min and stained according to the manufac-turer’s protocol. For PLA between overexpressed YES-YFP and FGFR3-Flag, RCS cells were transfected using FuGENE6 (1 g of DNA per well of a 24-well plate), and the PLA events were calcu-lated in three dimensions and normalized to YFP signal. Cells were analyzed by confocal microscopy (LSM700 laser scanning micro-scope, ZEISS) as described above.

Alcian blue stainingCells were fixed with 4% paraformaldehyde for 10 min, stained with 1% Alcian blue 8GX (Sigma-Aldrich), thoroughly washed with dis-tilled water, and mounted in glycerol. Brightfield images of Alcian blue samples (Fig. 5E and fig. S1A) were taken using the Olympus microscope IX71. Phase-contrast images were taken using a Carl ZEISS Axio Observer microscope. The contrast, brightness, and gamma were adjusted.

Densitometry and statistical analysisDensitometry was conducted using Gel Analyzer function in Fiji software. The optical density of bands was measured four times to obtain mean and SD values, which were then normalized. Statistical significance of differences was evaluated by Welch’s t test. Where appropriate, Bonferroni’s correction of P values was used.

SUPPLEMENTARY MATERIALSwww.sciencesignaling.org/cgi/content/full/11/548/eaap8608/DC1Fig. S1. Ship2 deletion or knockdown by CRISPR/Cas9 impairs Fgfr3-induced ECM degradation but not cell spreading in RCS cells.Fig. S2. Ship2 deletion or knockdown by CRISPR/Cas9 does not affect Fgfr2 and Fgfr3 abundance in RCS cells.Fig. S3. Deletion or knockdown of Ship2 impairs FGF2-mediated adaptor phosphorylation and activation of Erk.Fig. S4. Densitometric quantifications of Western blot analyses.Fig. S5. No substantial impairment of NGF- or EGF-induced changes in ERK activity in Ship2Crispr cells.Fig. S6. FGFR3 interacts with the SFKs FYN, LYN, FGR, and BLK.Fig. S7. FGR associates with both wild-type and catalytically inactive SHIP2.Fig. S8. Densitometric quantifications of Western blot analyses.Table S1. Expression vectors used in this study.Table S2. Antibodies used in this study.

REFERENCES AND NOTES 1. P. Krejci, The paradox of FGFR3 signaling in skeletal dysplasia: Why chondrocytes

growth arrest while other cells over proliferate. Mutat. Res. Rev. Mutat. Res. 759, 40–48 (2014).

2. D. M. Ornitz, P. J. Marie, Fibroblast growth factor signaling in skeletal development and disease. Genes Dev. 29, 1463–1486 (2015).

3. L. Bonafe, V. Cormier-Daire, C. Hall, R. Lachman, G. Mortier, S. Mundlos, G. Nishimura, L. Sangiorgi, R. Savarirayan, D. Sillence, J. Spranger, A. Superti-Furga, M. Warman, S. Unger, Nosology and classification of genetic skeletal disorders: 2015 revision. Am. J. Med. Genet. A 167A, 2869–2892 (2015).

4. J. Wesche, K. Haglund, E. M. Haugsten, Fibroblast growth factors and their receptors in cancer. Biochem. J. 437, 199–213 (2011).

5. P. Dvorak, D. Dvorakova, S. Koskova, M. Vodinska, M. Najvirtova, D. Krekac, A. Hampl, Expression and potential role of fibroblast growth factor 2 and its receptors in human embryonic stem cells. Stem Cells 23, 1200–1211 (2005).

6. J. Li, G. Wang, C. Wang, Y. Zhao, H. Zhang, Z. Tan, Z. Song, M. Ding, H. Deng, MEK/ERK signaling contributes to the maintenance of human embryonic stem cell self-renewal. Differentiation 75, 299–307 (2007).

7. O. Gafni, L. Weinberger, A. A. Mansour, Y. S. Manor, E. Chomsky, D. Ben-Yosef, Y. Kalma, S. Viukov, I. Maza, A. Zviran, Y. Rais, Z. Shipony, Z. Mukamel, V. Krupalnik, M. Zerbib, S. Geula, I. Caspi, D. Schneir, T. Shwartz, S. Gilad, D. Amann-Zalcenstein, S. Benjamin, I. Amit, A. Tanay, R. Massarwa, N. Novershtern, J. H. Hanna, Derivation of novel human ground state naive pluripotent stem cells. Nature 504, 282–286 (2013).

8. S. Kang, S. Dong, T.-L. Gu, A. Guo, M. S. Cohen, S. Lonial, H. J. Khoury, D. Fabbro, D. G. Gilliland, P. L. Bergsagel, J. Taunton, R. D. Polakiewicz, J. Chen, FGFR3 activates RSK2 to mediate hematopoietic transformation through tyrosine phosphorylation of RSK2 and activation of the MEK/ERK pathway. Cancer Cell 12, 201–214 (2007).

9. V. Yadav, X. Zhang, J. Liu, S. Estrem, S. Li, X.-Q. Gong, S. Buchanan, J. R. Henry, J. J. Starling, S.-B. Peng, Reactivation of mitogen-activated protein kinase (MAPK) pathway by FGF receptor 3 (FGFR3)/Ras mediates resistance to vemurafenib in human B-RAF V600E mutant melanoma. J. Biol. Chem. 287, 28087–28098 (2012).

10. H. Ishikawa, N. Tsuyama, S. Liu, S. Abroun, F.-J. Li, K.-I. Otsuyama, X. Zheng, Z. Ma, Y. Maki, M. S. Iqbal, M. Obata, M. M. Kawano, Accelerated proliferation of myeloma cells by interleukin-6 cooperating with fibroblast growth factor receptor 3-mediated signals. Oncogene 24, 6328–6332 (2005).

11. Y. Zhen, V. Sørensen, Y. Jin, Z. Suo, A. Wiedłocha, Indirubin-3′-monoxime inhibits autophosphorylation of FGFR1 and stimulates ERK1/2 activity via p38 MAPK. Oncogene 26, 6372–6385 (2007).

12. P. Krejci, J. Prochazkova, J. Smutny, K. Chlebova, P. Lin, A. Aklian, V. Bryja, A. Kozubik, W. R. Wilcox, FGFR3 signaling induces a reversible senescence phenotype in chondrocytes similar to oncogene-induced premature senescence. Bone 47, 102–110 (2010).

13. L. Dailey, E. Laplantine, R. Priore, C. Basilico, A network of transcriptional and signaling events is activated by FGF to induce chondrocyte growth arrest and differentiation. J. Cell Biol. 161, 1053–1066 (2003).

14. T. Matsushita, W. R. Wilcox, Y. Y. Chan, A. Kawanami, H. Bükülmez, G. Balmes, P. Krejci, P. B. Mekikian, K. Otani, I. Yamaura, M. L. Warman, D. Givol, S. Murakami, FGFR3 promotes synchondrosis closure and fusion of ossification centers through the MAPK pathway. Hum. Mol. Genet. 18, 227–240 (2009).

15. S. Murakami, G. Balmes, S. McKinney, Z. Zhang, D. Givol, B. de Crombrugghe, Constitutive activation of MEK1 in chondrocytes causes Stat1-independent achondroplasia-like dwarfism and rescues the Fgfr3-deficient mouse phenotype. Genes Dev. 18, 290–305 (2004).

16. H. Kouhara, Y. R. Hadari, T. Spivak-Kroizman, J. Schilling, D. Bar-Sagi, I. Lax, J. Schlessinger, A lipid-anchored Grb2-binding protein that links FGF-receptor activation to the Ras/MAPK signaling pathway. Cell 89, 693–702 (1997).

17. Y. R. Hadari, H. Kouhara, I. Lax, J. Schlessinger, Binding of Shp2 tyrosine phosphatase to FRS2 is essential for fibroblast growth factor-induced PC12 cell differentiation. Mol. Cell. Biol. 18, 3966–3973 (1998).

18. M. Buchtova, V. Oralova, A. Aklian, J. Masek, I. Vesela, Z. Ouyang, T. Obadalova, Z. Konecna, T. Spoustova, T. Pospisilova, P. Matula, M. Varecha, L. Balek, I. Gudernova, I. Jelinkova, I. Duran, I. Cervenkova, S. Murakami, A. Kozubik, P. Dvorak, V. Bryja, P. Krejci, Fibroblast growth factor and canonical WNT/-catenin signaling cooperate in suppression of chondrocyte differentiation in experimental models of FGFR signaling in cartilage. Biochim. Biophys. Acta 1852, 839–850 (2015).

19. L. Balek, P. Nemec, P. Konik, M. K. Bosakova, M. Varecha, I. Gudernova, J. Medalova, D. Krakow, P. Krejci, Proteomic analyses of signalling complexes associated with receptor tyrosine kinase identify novel members of fibroblast growth factor receptor 3 interactome. Cell. Signal. 42, 144–154 (2018).

20. P. L. Tavormina, R. Shiang, L. M. Thompson, Y.-Z. Zhu, D. J. Wilkin, R. S. Lachman, W. R. Wilcox, D. L. Rimoin, D. H. Cohn, J. J. Wasmuth, Thanatophoric dysplasia (types I and II) caused by distinct mutations in fibroblast growth factor receptor 3. Nat. Genet. 9, 321–328 (1995).

21. N. Prasad, R. S. Topping, S. J. Decker, SH2-containing inositol 5′-phosphatase SHIP2 associates with the p130Cas adapter protein and regulates cellular adhesion and spreading. Mol. Cell. Biol. 21, 1416–1428 (2001).

22. C. Erneux, S. Ghosh, A. R. Ramos, W. E. Edimo, New functions of the inositol polyphosphate 5-phosphatases in cancer. Curr. Pharm. Des. 22, 2309–2314 (2016).

on February 21, 2020

http://stke.sciencemag.org/

Dow

nloaded from

Page 13: CELL BIOLOGY Copyright © 2018 The inositol phosphatase ... · Fafilek et al., Sci. Signal. 11, eaap8608 2018 18 September 2018 SCIENCE SIGNALING| RESEARCH ARTICLE 1 of 14 CELL BIOLOGY

Fafilek et al., Sci. Signal. 11, eaap8608 (2018) 18 September 2018

S C I E N C E S I G N A L I N G | R E S E A R C H A R T I C L E

13 of 14

23. S. Ghosh, C. Huber, Q. Siour, S. B. Sousa, M. Wright, V. Cormier-Daire, C. Erneux, Fibroblasts derived from patients with opsismodysplasia display SHIP2-specific cell migration and adhesion defects. Hum. Mutat. 38, 1731–1739 (2017).

24. T. Aikawa, G. V. Segre, K. Lee, Fibroblast growth factor inhibits chondrocytic growth through induction of p21 and subsequent inactivation of cyclin E-Cdk2. J. Biol. Chem. 276, 29347–29352 (2001).

25. O. Rozenblatt-Rosen, E. Mosonego-Ornan, E. Sadot, L. Madar-Shapiro, Y. Sheinin, D. Ginsberg, A. Yayon, Induction of chondrocyte growth arrest by FGF: Transcriptional and cytoskeletal alterations. J. Cell Sci. 115 (Pt. 3), 553–562 (2002).

26. B. Fafilek, M. Hampl, N. Ricankova, I. Vesela, L. Balek, M. Kunova Bosakova, I. Gudernova, M. Varecha, M. Buchtova, P. Krejci, Statins do not inhibit the FGFR signaling in chondrocytes. Osteoarthr. Cartil. 25, 1522–1530 (2017).

27. P. Krejci, B. Masri, L. Salazar, C. Farrington-Rock, H. Prats, L. M. Thompson, W. R. Wilcox, Bisindolylmaleimide I suppresses fibroblast growth factor-mediated activation of Erk MAP kinase in chondrocytes by preventing Shp2 association with the Frs2 and Gab1 adaptor proteins. J. Biol. Chem. 282, 2929–2936 (2007).

28. I. Gudernova, S. Foldynova-Trantirkova, B. El Ghannamova, B. Fafilek, M. Varecha, L. Balek, E. Hruba, L. Jonatova, I. Jelinkova, M. K. Bosakova, L. Trantirek, J. Mayer, P. Krejci, One reporter for in-cell activity profiling of majority of protein kinase oncogenes. eLife 6, e21536 (2017).

29. H.-T. Lai, C.-M. Chiang, Bimolecular fluorescence complementation (BiFC) assay for direct visualization of protein-protein interaction in vivo. Bio Protoc. 3, e935 (2013).

30. T. Wada, T. Sasaoka, M. Funaki, H. Hori, S. Murakami, M. Ishiki, T. Haruta, T. Asano, W. Ogawa, H. Ishihara, M. Kobayashi, Overexpression of SH2-containing inositol phosphatase 2 results in negative regulation of insulin-induced metabolic actions in 3T3-L1 adipocytes via Its 5′-phosphatase catalytic activity. Mol. Cell. Biol. 21, 1633–1646 (2001).

31. Y. Ichihara, R. Fujimura, H. Tsuneki, T. Wada, K. Okamoto, H. Gouda, S. Hirono, K. Sugimoto, Y. Matsuya, T. Sasaoka, N. Toyooka, Rational design and synthesis of 4-substituted 2-pyridin-2-ylamides with inhibitory effects on SH2 domain-containing inositol 5′-phosphatase 2 (SHIP2). Eur. J. Med. Chem. 62, 649–660 (2013).

32. A. Raucci, E. Laplantine, A. Mansukhani, C. Basilico, Activation of the ERK1/2 and p38 mitogen-activated protein kinase pathways mediates fibroblast growth factor-induced growth arrest of chondrocytes. J. Biol. Chem. 279, 1747–1756 (2004).

33. C. J. Marshall, Specificity of receptor tyrosine kinase signaling: Transient versus sustained extracellular signal-regulated kinase activation. Cell 80, 179–185 (1995).

34. M. A. Lemmon, J. Schlessinger, Cell signaling by receptor tyrosine kinases. Cell 141, 1117–1134 (2010).

35. S. Yamada, T. Taketomi, A. Yoshimura, Model analysis of difference between EGF pathway and FGF pathway. Biochem. Biophys. Res. Commun. 314, 1113–1120 (2004).

36. P. Krejci, V. Bryja, J. Pachernik, A. Hampl, R. Pogue, P. Mekikian, W. R. Wilcox, FGF2 inhibits proliferation and alters the cartilage-like phenotype of RCS cells. Exp. Cell Res. 297, 152–164 (2004).

37. V. Knights, S. J. Cook, De-regulated FGF receptors as therapeutic targets in cancer. Pharmacol. Ther. 125, 105–117 (2010).

38. I. Mellman, Y. Yarden, Endocytosis and cancer. Cold Spring Harb. Perspect. Biol. 5, a016949 (2013).

39. B. Nadratowska-Wesolowska, E. M. Haugsten, M. Zakrzewska, P. Jakimowicz, Y. Zhen, D. Pajdzik, J. Wesche, A. Wiedlocha, RSK2 regulates endocytosis of FGF receptor 1 by phosphorylation on serine 789. Oncogene 33, 4823–4836 (2014).

40. M. Chesi, E. Nardini, L. A. Brents, E. Schröck, T. Ried, W. M. Kuehl, P. L. Bergsagel, Frequent translocation t(4;14)(p16.3;q32.3) in multiple myeloma is associated with increased expression and activating mutations of fibroblast growth factor receptor 3. Nat. Genet. 16, 260–264 (1997).

41. L. Xie, X. Su, L. Zhang, X. Yin, L. Tang, X. Zhang, Y. Xu, Z. Gao, K. Liu, M. Zhou, B. Gao, D. Shen, L. Zhang, J. Ji, P. R. Gavine, J. Zhang, E. Kilgour, X. Zhang, Q. Ji, FGFR2 gene amplification in gastric cancer predicts sensitivity to the selective FGFR inhibitor AZD4547. Clin. Cancer Res. 19, 2572–2583 (2013).

42. J. Weiss, M. L. Sos, D. Seidel, M. Peifer, T. Zander, J. M. Heuckmann, R. T. Ullrich, R. Menon, S. Maier, A. Soltermann, H. Moch, P. Wagener, F. Fischer, S. Heynck, M. Koker, J. Schöttle, F. Leenders, F. Gabler, I. Dabow, S. Querings, L. C. Heukamp, H. Balke-Want, S. Ansén, D. Rauh, I. Baessmann, J. Altmüller, Z. Wainer, M. Conron, G. Wright, P. Russell, B. Solomon, E. Brambilla, C. Brambilla, P. Lorimier, S. Sollberg, O. T. Brustugun, W. Engel-Riedel, C. Ludwig, I. Petersen, J. Sänger, J. Clement, H. Groen, W. Timens, H. Sietsma, E. Thunnissen, E. Smit, D. Heideman, F. Cappuzzo, C. Ligorio, S. Damiani, M. Hallek, R. Beroukhim, W. Pao, B. Klebl, M. Baumann, R. Buettner, K. Ernestus, E. Stoelben, J. Wolf, P. Nürnberg, S. Perner, R. K. Thomas, Frequent and focal FGFR1 amplification associates with therapeutically tractable FGFR1 dependency in squamous cell lung cancer. Sci. Transl. Med. 2, 62ra93 (2010).

43. A. Wong, B. Lamothe, A. Lee, J. Schlessinger, I. Lax, A. Li, FRS2 attenuates FGF receptor signaling by Grb2-mediated recruitment of the ubiquitin ligase Cbl. Proc. Natl. Acad. Sci. U.S.A. 99, 6684–6689 (2002).

44. I. Lax, A. Wong, B. Lamothe, A. Lee, A. Frost, J. Hawes, J. Schlessinger, The docking protein FRS2 controls a MAP kinase-mediated negative feedback mechanism for signaling by FGF receptors. Mol. Cell 10, 709–719 (2002).

45. N. Gotoh, Regulation of growth factor signaling by FRS2 family docking/scaffold adaptor proteins. Cancer Sci. 99, 1319–1325 (2008).

46. M. Muda, U. Boschert, R. Dickinson, J.-C. Martinou, I. Martinou, M. Camps, W. Schlegel, S. Arkinstall, MKP-3, a novel cytosolic protein-tyrosine phosphatase that exemplifies a new class of mitogen-activated protein kinase phosphatase. J. Biol. Chem. 271, 4319–4326 (1996).

47. A. Sasaki, T. Taketomi, R. Kato, K. Saeki, A. Nonami, M. Sasaki, M. Kuriyama, N. Saito, M. Shibuya, A. Yoshimura, Mammalian Sprouty4 suppresses Ras-independent ERK activation by binding to Raf1. Nat. Cell Biol. 5, 427–432 (2003).

48. N. E. Kremer, G. D’Arcangelo, S. M. Thomas, M. DeMarco, J. S. Brugge, S. Halegoua, Signal transduction by nerve growth factor and fibroblast growth factor in PC12 cells requires a sequence of src and ras actions. J. Cell Biol. 115, 809–819 (1991).

49. J. M. Rodier, A. M. Vallés, M. Denoyelle, J. P. Thiery, B. Boyer, pp60c-src is a positive regulator of growth factor-induced cell scattering in a rat bladder carcinoma cell line. J. Cell Biol. 131, 761–773 (1995).

50. A. P. Belsches, M. D. Haskell, S. J. Parsons, Role of c-Src tyrosine kinase in EGF-induced mitogenesis. Front. Biosci. 2, d501–d518 (1997).

51. J. Liu, C. Huang, X. Zhan, Src is required for cell migration and shape changes induced by fibroblast growth factor 1. Oncogene 18, 6700–6706 (1999).

52. P. Klint, S. Kanda, Y. Kloog, L. Claesson-Welsh, Contribution of Src and Ras pathways in FGF-2 induced endothelial cell differentiation. Oncogene 18, 3354–3364 (1999).

53. F. Debiais, J. Lemonnier, E. Hay, P. Delannoy, J. Caverzasio, P. J. Marie, Fibroblast growth factor-2 (FGF-2) increases N-cadherin expression through protein kinase C and Src-kinase pathways in human calvaria osteoblasts. J. Cell. Biochem. 81, 68–81 (2001).

54. D. M. Kilkenny, J. V Rocheleau, J. Price, M. B. Reich, G. G. Miller, c-Src regulation of fibroblast growth factor-induced proliferation in murine embryonic fibroblasts. J. Biol. Chem. 278, 17448–17454 (2003).

55. S. Knuchel, P. Anderle, P. Werfelli, E. Diamantis, C. Rüegg, Fibroblast surface-associated FGF-2 promotes contact-dependent colorectal cancer cell migration and invasion through FGFR-SRC signaling and integrin v5-mediated adhesion. Oncotarget 6, 14300–14317 (2015).

56. X. Zhan, C. Plourde, X. Hu, R. Friesel, T. Maciag, Association of fibroblast growth factor receptor-1 with c-Src correlates with association between c-Src and cortactin. J. Biol. Chem. 269, 20221–20224 (1994).

57. J. Hama, H. Xu, M. Goldfarb, D. C. Weinstein, SNT-1/FRS2 physically interacts with Laloo and mediates mesoderm induction by fibroblast growth factor. Mech. Dev. 109, 195–204 (2001).

58. P. Zhang, J. S. S. Greendorfer, J. Jiao, S. C. C. Kelpke, J. A. A. Thompson, Alternatively spliced FGFR-1 isoforms differentially modulate endothelial cell activation of c-YES. Arch. Biochem. Biophys. 450, 50–62 (2006).

59. T. Shono, H. Kanetake, S. Kanda, The role of mitogen-activated protein kinase activation within focal adhesions in chemotaxis toward FGF-2 by murine brain capillary endothelial cells. Exp. Cell Res. 264, 275–283 (2001).

60. S. Clément, U. Krause, F. Desmedt, J.-F. Tanti, J. Behrends, X. Pesesse, T. Sasaki, J. Penninger, M. Doherty, W. Malaisse, J. E. Dumont, Y. Le Marchand-Brustel, C. Erneux, L. Hue, S. Schurmans, The lipid phosphatase SHIP2 controls insulin sensitivity. Nature 409, 92–97 (2001).

61. M. W. Sleeman, K. E. Wortley, K.-M. V. Lai, L. C. Gowen, J. Kintner, W. O. Kline, K. Garcia, T. N. Stitt, G. D. Yancopoulos, S. J. Wiegand, D. J. Glass, Absence of the lipid phosphatase SHIP2 confers resistance to dietary obesity. Nat. Med. 11, 199–205 (2005).

62. B. Li, D. Krakow, D. A. Nickerson, M. J. Bamshad; University of Washington Center for Mendelian Genomics, Y. Chang, R. S. Lachman, A. Yilmaz, H. Kayserili, D. H. Cohn, Opsismodysplasia resulting from an insertion mutation in the SH2 domain, which destabilizes INPPL1. Am. J. Med. Genet. A 164A, 2407–2411 (2014).

63. J. E. Below, D. L. Earl, K. M. Shively, M. J. McMillin, J. D. Smith, E. H. Turner, M. J. Stephan, L. I. Al-Gazali, J. L. Hertecant, D. Chitayat, S. Unger, D. H. Cohn, D. Krakow, J. M. Swanson, E. M. Faustman, J. Shendure, D. A. Nickerson, M. J. Bamshad; University of Washington Center for Mendelian Genomics, Whole-genome analysis reveals that mutations in inositol polyphosphate phosphatase-like 1 cause opsismodysplasia. Am. J. Hum. Genet. 92, 137–143 (2013).

64. C. Feist, P. Holden, J. Fitzgerald, Novel compound heterozygous mutations in inositol polyphosphate phosphatase-like 1 in a family with severe opsismodysplasia. Clin. Dysmorphol. 25, 152–155 (2016).

65. H. Lee, L. Nevarez, R. S. Lachman, W. R. Wilcox, D. Krakow, D. H. Cohn; University of Washington Center for Mendelian Genomics, A second locus for Schneckenbecken dysplasia identified by a mutation in the gene encoding inositol polyphosphate phosphatase-like 1 (INPPL1). Am. J. Med. Genet. A 167A, 2470–2473 (2015).

66. V. Cormier-Daire, A. L. Delezoide, N. Philip, P. Marcorelles, K. Casas, Y. Hillion, L. Faivre, D. L. Rimoin, A. Munnich, P. Maroteaux, M. Le Merrer, Clinical, radiological, and

on February 21, 2020

http://stke.sciencemag.org/

Dow

nloaded from

Page 14: CELL BIOLOGY Copyright © 2018 The inositol phosphatase ... · Fafilek et al., Sci. Signal. 11, eaap8608 2018 18 September 2018 SCIENCE SIGNALING| RESEARCH ARTICLE 1 of 14 CELL BIOLOGY

Fafilek et al., Sci. Signal. 11, eaap8608 (2018) 18 September 2018

S C I E N C E S I G N A L I N G | R E S E A R C H A R T I C L E

14 of 14

chondro-osseous findings in opsismodysplasia: Survey of a series of 12 unreported cases. J. Med. Genet. 40, 195–200 (2003).

67. C. Fyffe, M. Falasca, 3-Phosphoinositide-dependent protein kinase-1 as an emerging target in the management of breast cancer. Cancer Manag. Res. 5, 271–280 (2013).

68. C. Huber, E. A. Faqeih, D. Bartholdi, C. Bole-Feysot, Z. Borochowitz, D. P. Cavalcanti, A. Frigo, P. Nitschke, J. Roume, H. G. Santos, S. A. Shalev, A. Superti-Furga, A.-L. Delezoide, M. Le Merrer, A. Munnich, V. Cormier-Daire, Exome sequencing identifies INPPL1 mutations as a cause of opsismodysplasia. Am. J. Hum. Genet. 92, 144–149 (2013).

69. A. Fradet, J. Fitzgerald, INPPL1 gene mutations in opsismodysplasia. J. Hum. Genet. 62, 135–140 (2017).

70. A. Kasprzak, W. Kwasniewski, A. Adamek, A. Gozdzicka-Jozefiak, Insulin-like growth factor (IGF) axis in cancerogenesis. Mutat. Res. Rev. Mutat. Res. 772, 78–104 (2017).

71. J. Baker, J. P. Liu, E. J. Robertson, A. Efstratiadis, Role of insulin-like growth factors in embryonic and postnatal growth. Cell 75, 73–82 (1993).

72. Y. Wang, S. Nishida, T. Sakata, H. Z. Elalieh, W. Chang, B. P. Halloran, S. B. Doty, D. D. Bikle, Insulin-like growth factor-I is essential for embryonic bone development. Endocrinology 147, 4753–4761 (2006).

73. Y. Wang, Z. Cheng, H. Z. Elalieh, E. Nakamura, M.-T. Nguyen, S. Mackem, T. L. Clemens, D. D. Bikle, W. Chang, IGF-1R signaling in chondrocytes modulates growth plate development by interacting with the PTHrP/Ihh pathway. J. Bone Miner. Res. 26, 1437–1446 (2011).

74. K. Mukhopadhyay, V. Lefebvre, G. Zhou, S. Garofalo, J. H. Kimura, B. de Crombrugghe, Use of a new rat chondrosarcoma cell line to delineate a 119-base pair chondrocyte-specific enhancer element and to define active promoter segments in the mouse pro-1(II) collagen gene. J. Biol. Chem. 270, 27711–27719 (1995).

75. E. M. Haugsten, A. Brech, K. Liestøl, J. C. Norman, J. Wesche, Photoactivation approaches reveal a role for Rab11 in FGFR4 recycling and signalling. Traffic 15, 665–683 (2014).

76. M. Kostas, E. M. Haugsten, Y. Zhen, V. Sorensen, P. Szybowska, E. Fiorito, S. Lorenz, G. A. de Souza, A. Wiedlocha, J. Wesche, Protein tyrosine phosphatase receptor type G (PTPRG) controls fibroblast growth factor receptor (FGFR) 1 activity and influences sensitivity to FGFR kinase inhibitors. Mol. Cell. Proteomics 17, 850–870 (2018).

77. W. Elong Edimo, S. Ghosh, R. Derua, V. Janssens, E. Waelkens, J.-M. Vanderwinden, P. Robe, C. Erneux, SHIP2 controls plasma membrane PI(4,5)P2 thereby participating in the control of cell migration in 1321 N1 glioblastoma cells. J. Cell Sci. 129, 1101–1114 (2016).

78. I. Gudernova, I. Vesela, L. Balek, M. Buchtova, H. Dosedelova, M. Kunova, J. Pivnicka, I. Jelinkova, L. Roubalova, A. Kozubik, P. Krejci, Multikinase activity of fibroblast growth factor receptor (FGFR) inhibitors SU5402, PD173074, AZD1480, AZD4547 and BGJ398 compromises the use of small chemicals targeting FGFR catalytic activity for therapy of short-stature syndromes. Hum. Mol. Genet. 25, 9–23 (2016).

79. F. A. Ran, P. D. Hsu, J. Wright, V. Agarwala, D. A. Scott, F. Zhang, Genome engineering using the CRISPR-Cas9 system. Nat. Protoc. 8, 2281–2308 (2013).

80. T. G. Montague, J. M. Cruz, J. A. Gagnon, G. M. Church, E. Valen, CHOPCHOP: A CRISPR/Cas9 and TALEN web tool for genome editing. Nucleic Acids Res. 42, W401–W407 (2014).

81. Y. R. Hadari, N. Gotoh, H. Kouhara, I. Lax, J. Schlessinger, Critical role for the docking-protein FRS2 in FGF receptor-mediated signal transduction pathways. Proc. Natl. Acad. Sci. U.S.A. 98, 8578–8583 (2001).

Acknowledgments: We thank P. Nemec for excellent technical assistance, V. Korinek for the GReP plasmid, and A. Wiedlocha for helpful discussions. Funding: This study was supported by Ministry of Education, Youth and Sports of the Czech Republic (KONTAKT II LH15231), Agency for Healthcare Research of the Czech Republic (15-33232A, 15-34405A, and NV18-08-00567), Czech Science Foundation (GA17-09525S), National Program of Sustainability II [MEYS CR (LQ1605) and CEITEC 2020 (LQ1601)]. This study was partially supported by the Research Council of Norway through its Centers of Excellence funding scheme (project number 262652). E.M.H. has a Career fellowship (project number 6842225) from the Norwegian Cancer Society. M.K.B. was supported by junior researcher funds from the Faculty of Medicine, Masaryk University. J.T.Z. was supported by Kohn fellowship at Masaryk University and National Institute of Arthritis and Musculoskeletal and Skin Diseases of the NIH (1F31AR066487). D.K. is supported by NIH grants R01-AR066124 and R01-AR062651. The content is solely the responsibility of the authors and does not necessarily represent the official views of the NIH. C.E. is supported by grants from the Université Libre de Bruxelles and of the Fonds de la Recherche Scientifique Médicale (J.0078.18). S.G. is supported by Hoguet, Fonds Lekime-Ropsy, and Télévie fellowships. Author contributions: P.K. designed the project. C.E., L.T., D.K., J.W., and P.K. designed experiments. B.F., L.B., M.K.B., M.V., A.N., T.G., I.G., J.K., S.G., M.P., L.J., N.H.C., J.T.Z., M.K., E.M.H., and P.K. performed experiments. B.F., M.K.B., C.E., J.W., D.K., and P.K. wrote the manuscript. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper or the Supplementary Materials.

Submitted 4 September 2017Accepted 29 August 2018Published 18 September 201810.1126/scisignal.aap8608

Citation: B. Fafilek, L. Balek, M. K. Bosakova, M. Varecha, A. Nita, T. Gregor, I. Gudernova, J. Krenova, S. Ghosh, M. Piskacek, L. Jonatova, N. H. Cernohorsky, J. T. Zieba, M. Kostas, E. M. Haugsten, J. Wesche, C. Erneux, L. Trantirek, D. Krakow, P. Krejci, The inositol phosphatase SHIP2 enables sustained ERK activation downstream of FGF receptors by recruiting Src kinases. Sci. Signal. 11, eaap8608 (2018).

on February 21, 2020

http://stke.sciencemag.org/

Dow

nloaded from

Page 15: CELL BIOLOGY Copyright © 2018 The inositol phosphatase ... · Fafilek et al., Sci. Signal. 11, eaap8608 2018 18 September 2018 SCIENCE SIGNALING| RESEARCH ARTICLE 1 of 14 CELL BIOLOGY

receptors by recruiting Src kinasesThe inositol phosphatase SHIP2 enables sustained ERK activation downstream of FGF

Ellen Margrethe Haugsten, Jørgen Wesche, Christophe Erneux, Lukas Trantirek, Deborah Krakow and Pavel KrejciJitka Krenova, Somadri Ghosh, Martin Piskacek, Lucie Jonatova, Nicole H. Cernohorsky, Jennifer T. Zieba, Michal Kostas, Bohumil Fafilek, Lukas Balek, Michaela Kunova Bosakova, Miroslav Varecha, Alexandru Nita, Tomas Gregor, Iva Gudernova,

DOI: 10.1126/scisignal.aap8608 (548), eaap8608.11Sci. Signal. 

pathologies.associated with oncogenesis and developmental disorders, SHIP2 may be a potential therapeutic target for thesethat mediated signal relay from FGFRs to ERK. Because sustained ERK activation due to aberrant FGFR signaling is scaffold that recruited Src family kinases to FGFR complexes, thus enhancing the phosphorylation of adaptor proteinsactivation into sustained ERK signaling. The catalytic activity of SHIP2 was not required. Instead, SHIP2 acted as a

. found that the inositol phosphatase SHIP2 was required for converting transient FGFRet alERK signaling. Fafilek receptors are endocytosed and degraded after activation. Nevertheless, FGFRs stimulate both sustained and transient

Activation of fibroblast growth factor receptors (FGFRs) stimulates downstream signaling transiently because theConverting transient to sustained signaling

ARTICLE TOOLS http://stke.sciencemag.org/content/11/548/eaap8608

MATERIALSSUPPLEMENTARY http://stke.sciencemag.org/content/suppl/2018/09/14/11.548.eaap8608.DC1

CONTENTRELATED

http://stke.sciencemag.org/content/sigtrans/12/585/eaay2985.fullhttp://stke.sciencemag.org/content/sigtrans/12/565/eaaw6857.fullhttp://stke.sciencemag.org/content/sigtrans/11/553/eaat5916.fullhttp://stke.sciencemag.org/content/sigtrans/9/450/rs12.fullhttp://stke.sciencemag.org/content/sigtrans/11/539/eaas9609.full

REFERENCES

http://stke.sciencemag.org/content/11/548/eaap8608#BIBLThis article cites 81 articles, 25 of which you can access for free

PERMISSIONS http://www.sciencemag.org/help/reprints-and-permissions

Terms of ServiceUse of this article is subject to the

is a registered trademark of AAAS.Science SignalingYork Avenue NW, Washington, DC 20005. The title (ISSN 1937-9145) is published by the American Association for the Advancement of Science, 1200 NewScience Signaling

Science. No claim to original U.S. Government WorksCopyright © 2018 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of

on February 21, 2020

http://stke.sciencemag.org/

Dow

nloaded from