Signal transduction by vascular endothelial growth factor receptors

46 downloads 243 Views 775KB Size Report
VEGFs (vascular endothelial growth factors) control vascular ... Therapeutic agents ... receptors and have therefore been described as anti-angiogenic. [8].
Biochem. J. (2011) 437, 169–183 (Printed in Great Britain)

169

doi:10.1042/BJ20110301

REVIEW ARTICLE

Signal transduction by vascular endothelial growth factor receptors Sina KOCH, S`onia TUGUES, Xiujuan LI, Laura GUALANDI and Lena CLAESSON-WELSH1 Department of Immunology, Genetics and Pathology, Uppsala University, Rudbeck Laboratory, Dag Hammarskj¨oldsv. 20, 751 85 Uppsala, Sweden

VEGFs (vascular endothelial growth factors) control vascular development during embryogenesis and the function of blood vessels and lymphatic vessels in the adult. There are five related mammalian ligands, which act through three receptor tyrosine kinases. Signalling is modulated through neuropilins, which act as VEGF co-receptors. Heparan sulfate and integrins are also important modulators of VEGF signalling. Therapeutic agents that interfere with VEGF signalling have been developed with

the aim of decreasing angiogenesis in diseases that involve tissue growth and inflammation, such as cancer. The present review will outline the current understanding and consequent biology of VEGF receptor signalling.

THE VEGF (VASCULAR ENDOTHELIAL GROWTH FACTOR) FAMILY

Several of the VEGF family of ligands and receptors, notably VEGFA, are regulated by HIF (hypoxia-inducible factor) [11], leading to increased expression during tissue growth both in health (wound healing and embryonic development) and disease (cancer). In addition, a wide range of other transcription factors and metabolic regulators, such as the ETS (gene transduced by the leukaemia virus E26) family of transcription factors [12] and ROS (reactive oxygen species) [13], modulate the expression levels of VEGFs as well as of the VEGF receptors. The VEGF ligands, in particular VEGFA, are produced by most parenchymal cells and act in a paracrine manner on adjacent endothelial cells to regulate VEGF receptor signalling and biology. Autocrine VEGFA may be essential for endothelial cell survival as shown in recombinant mice specifically lacking endothelium-produced VEGFA [14].

VEGF denotes a family of five related mammalian growth factors: VEGFA (the prototype), VEGFB, VEGFC, VEGFD and PlGF (placental growth factor). An alternative designation for VEGFA is VFP (vascular permeability factor) [1]. The VEGFs are homodimeric polypeptides, although naturally occurring heterodimers of VEGFA and PlGF have been described [2]. The complexity of the VEGF family is increased further by alternative splicing of VEGFA, VEGFB and PlGF, and processing of VEGFC and VEGFD [3]. VEGFA can also be processed by matrix metalloproteinases [4]. Splicing and processing regulates the ability of the ligands to bind to VEGF receptors, to HS (heparan sulfate) and NRPs (neuropilins) [5,6]. For example, VEGFA is alternatively spliced to generate VEGFA121, VEGFA145, VEGFA165 and VEGFA189 (the numbers indicating amino acid residues in the human polypeptides). VEGFA121 does not bind to HS and is freely diffusible, whereas VEGFA165 and VEGFA189 bind to HS as well as to NRP1. The restricted diffusion of HS-binding VEGFs affects the temporal and spatial distribution of VEGF receptor signalling and, in addition, it modulates signalling in a qualitative manner, e.g. by recruitment of integrins into the signalling complex [7]. Additional mammalian VEGF variants include a series of VEGFA splice variants denoted VEGFAxxxb, which bind, but fail to efficiently activate, VEGF receptors and have therefore been described as anti-angiogenic [8]. Furthermore, a number of non-vertebrate polypeptides with structural and functional relatedness to the mammalian VEGFs have been identified. These include the parapox virus open reading frame denoted VEGFE [9] and the snake venom-derived polypeptides denoted VEGFF [10].

Key words: angiogenesis, internalization, neuropilin, signal transduction, vascular endothelial growth factor (VEGF), vascular endothelial growth factor receptor (VEGFR).

THE VEGF RECEPTOR TYROSINE KINASE FAMILY

In humans, mice and other mammals, three structurally related VEGFRs (VEGF receptor tyrosine kinases) have been identified, namely VEGFR1, VEGFR2 and VEGFR3 (Figure 1). In the zebrafish, there are four genes encoding VEGFRs: the Flt (Fmslike tyrosine kinase) 1 orthologue (corresponding to VEGFR1), the Flt4 orthologue (corresponding to VEGFR3) and two genes with highest similarity to VEGFR2 denoted Kdr (kinase insert domain receptor)-like and Kdr [15]. The VEGFRs show a similar organization with an extracellular, ligand-binding, domain composed of immunoglobulin-like loops, a transmembrane domain, a juxtamembrane domain, a split tyrosine kinase domain

Abbreviations used: ARF, ADP-ribosylation factor; Bcl, B-cell lymphoma; BAD, Bcl-2-associated death promoter; CBL, Cas-Br-M murine ecotropic retroviral transforming sequence homologue; cPLA2 , cytoplasmic phospholipase A2 ; CREB, rcAMP-response-element-binding protein; CRK, C10 regulator of kinase; E, embryonic day; eNOS, endothelial nitric oxide synthase; ERK, extracellular-signal-regulated kinase; FAK, focal adhesion kinase; Flt, Fms-like tyrosine kinase; GAB, GRB2-binding protein; GRB, growth-factor-receptor-bound protein; H, heparin; HDAC, histone deacetylase; HS, heparan sulfate; HSP, heat-shock protein; HSPG, HS proteoglycan; IAP, inhibitor(s) of apoptosis; IQGAP1, IQ-motif-containing GTPase-activating protein 1; JNK, c-Jun N-terminal kinase; KDR, kinase insert domain receptor; MAPK, mitogen-activated protein kinase; Nck, non-catalytic region of tyrosine kinase adaptor protein; NRP, neuropilin; PAK, p21-activated protein kinase; PDK, phosphoinositide-dependent kinase; PI3K, phosphoinositide 3-kinase; PKC, protein kinase C; PLC, phospholipase C; PlGF, placental growth factor; PTP, protein tyrosine phosphatase; Rab, Ras-related GTPase; sFlt, soluble Flt; SH2, Src homology 2; SHB, SH2-domain-containing adaptor protein B; SHC, Src homology and collagen homology; Sck, SHC-related adaptor protein; TK, tyrosine kinase; TSAd, T-cell-specific adapter molecule; VE-cadherin, vascular endothelial cadherin; VEGF, vascular endothelial growth factor; hVEGF, human VEGF; VEGFR, VEGF receptor tyrosine kinase; sVEGFR, soluble VEGFR; VVO, vesiculo-vacuolar organelle. 1 To whom correspondence should be addressed (email [email protected]).  c The Authors Journal compilation  c 2011 Biochemical Society

170

Figure 1

S. Koch and others

VEGF binding specificities and VEGFR signalling complexes

Schematic outline of the five VEGFs, VEGFA, VEGFB, VEGFC, VEGFD and PlGF, binding with different affinities to three VEGFRs, initiating VEGFR homo- and hetero-dimer formation. Proteolytic processing of VEGFC and VEGFD allows binding to VEGFR2. VEGFR Ig-like domains involved in VEGF binding are indicated by hatched circles. Soluble VEGFRs (sVEGFR1 and sVEGFR2) lack the seventh Ig-like domain. It remains to be shown whether sVEGFRs occur as monomers, homodimers or heterodimers with full-length VEGFRs. JMD, juxtamembrane domain; KID, kinase insert domain; TMD, transmembrane domain; TKD1, ATP-binding domain; TKD2, phosphotransferase domain.

and a C-terminal tail (Figure 1). There is an overall pattern of VEGFR1 expression in monocytes and macrophages, VEGFR2 in vascular endothelial cells and VEGFR3 in lymphatic endothelial cells. However, with increasing sensitivity and accuracy of reagents and techniques, many exceptions to this basic pattern have been reported.

enhanced phosphatase 1), VEPTP (vascular endothelial PTP), SH2-domain PTP and PTP1B [22].

VEGFR1 SIGNALLING Structure and expression of VEGFR1

VEGFR DIMERIZATION AND ACTIVATION OF THE KINASE

Binding of VEGF occurs to the N-terminal part of the extracellular domain of the receptor (more detailed information is given in each VEGFR section below). Binding of VEGF to its VEGFR can occur in cis, e.g. by freely diffusible VEGF or by presentation of VEGF through co-receptors expressed on the same cell as the VEGFR, or in trans, e.g. by presentation through co-receptors expressed on adjacent cells (see the ‘VEGF co-receptors’ section below). Dimerization of receptors to form homo- or hetero-dimers [16] is necessary, but not sufficient, for receptor activation [17,18]. Dimerization is stabilized by contact points between receptor extracellular domains in Ig-loop 7 [19,20]. The orientation of the receptor monomers is influenced by the transmembrane domains [17]. The resulting rigid arrangement of two receptor monomers is required for exact positioning of the intracellular kinase domains [19]. According to the consensus model for activation of receptor tyrosine kinases [21], dimerization is accompanied by changes in the intracellular domain conformation. Formal proof is missing, however, due to challenges in structural analyses of large transmembrane proteins. The conformational changes in turn lead to exposure of the ATP-binding site in the intracellular kinase domain, followed by binding of ATP and auto- or transphosphorylation of tyrosine residues on the receptor dimer itself as well as on downstream signal transducers, often SH2 (Src homology 2)-domain-containing proteins. Whether a particular tyrosine residue will serve as a substrate for the inherent kinase cannot be predicted, and the identification of phosphotyrosine sites in the receptors rely on in vitro analyses. However, as discussed below, studies of mice carrying mutations in VEGFR2 have begun to shed light on in vivo functions of specific phosphorylation sites. Tyrosine phosphorylation is tightly regulated by internalization and degradation (see below) or by dephosphorylation through PTPs (protein tyrosine phosphatases), such as DEP1 (density c The Authors Journal compilation  c 2011 Biochemical Society

VEGFR1 (also called Flt1 in the mouse) is a 180–185 kDa glycoprotein [23,24] that is activated in response to binding of VEGFA, VEGFB and PlGF (Figure 1). The ligands bind to Igloop 2, but loops 1 and 3 are also required for high-affinity binding [25,26]. The crystal structure of the receptor-binding domain of VEGF bound to the second Ig-loop of VEGFR1 shows hydrophobic interactions stabilizing the ligand–receptor dimers [27]. There is a similar pattern for PlGF binding to VEGFR1 [28]. VEGFR1 is expressed in vascular endothelial cells at relatively high levels throughout development and in the adult [29,30]. In addition, a wide range of non-endothelial cells, such as monocytes and macrophages, human trophoblasts, renal mesangial cells, vascular smooth muscle cells, dendritic cells and different human tumour cell types, express VEGFR1 [31–35]. VEGFR1 expression is regulated by hypoxia through a hypoxia-inducible enhancer element in the VEGFR1 promoter [36]. Alternative splicing of VEGFR1 results in the generation of sVEGFR (soluble VEGFR) 1 [also denoted sFlt1 (soluble Flt1)], encompassing the N-terminal six extracellular Ig-loops [37]. VEGFR1 splicing is controlled by intronic signals [38], and gives rise to abundant expression of sFlt1 in the placenta (see below). Truncated intracellular variants of VEGFR1 have also been described [39].

VEGFR1 signal transduction

Although VEGFR1 binds VEGFA with higher affinity than does VEGFR2 (K d = 15 pM compared with 750 pM) [40], VEGFR1 tyrosine kinase activity is only weakly induced by its ligands; several underlying mechanisms have been suggested. First, Gille et al. [41] identified a repressor sequence in the juxtamembrane domain of VEGFR1. Secondly, structural properties of the activation loop of VEGFR1 [42], including the lack of positive regulatory tyrosine residues [43], contribute to the poor kinase activity. By overexpression of VEGFR1 in insect cells or mammalian cells, several VEGFR1 tyrosine phosphorylation sites have been identified, namely Tyr794 , Tyr1169 , Tyr1213 , Tyr1242 ,

Signal transduction by VEGF receptors

Figure 2

VEGFR1 tyrosine phosphorylation sites and signal transduction

Schematic outline of activated and dimerized VEGFR1 with tyrosine (Y) phosphorylation sites (indicated by numbers) and downstream signalling pathways. Hatched circles represent ligand-binding domains. Tyr1309 (boxed italics) is phosphorylated only upon PlGF stimulation. A repressor sequence (purple) in the juxtamembrane domain is one possible mechanism for the weak kinase activity of VEGFR1, another being the lack of phosphorylation of tyrosine residues in the kinase activation loop. Binding of signalling mediators (‘rocket’ shapes) to certain phosphorylation sites initiates activation of downstream signalling molecules (ovals), leading to specific biological responses (boxes). Certain signalling pathways are not yet reported to connect to a particular phosphotyrosine site in VEGFR1 and detailed information on signalling pathways may be lacking (broken arrows). Biological functions of VEGFR1 are summarized in the bottom box. See the text for details. 3D, three-dimensional; JAK, Janus kinase; NFAT, nuclear factor of activated T-cells; PKB, protein kinase B; RACK1, receptor for activated C-kinase 1; SHP2, SH2-domain-containing protein tyrosine phosphatase 2; STAT, signal transducer and activator of transcription.

Tyr1327 and Tyr1333 [43–45]. Interestingly, the pattern of tyrosine phosphorylation of VEGFR1 depends on the ligand, i.e. PlGF, but not VEGFA, induces phosphorylation of Tyr1309 [46] (Figure 2). A range of signalling molecules associate with VEGFR1 phosphorylation sites in vitro. Tyr794 [44] and Tyr1169 [47] are involved in binding and activation of PLC (phospholipase C) γ , resulting in intracellular Ca2 + fluxes and generation of inositol 1,4,5-trisphosphate [47]. Tyr1213 binds different SH2containing proteins, such as PLCγ , GRB (growth-factor-receptorbound protein) 2, Nck (non-catalytic region of tyrosine kinase adaptor protein) and SHP-2 (SH2-domain-containing protein tyrosine phosphatase 2) [43,48], and, according to other studies, also the p85 subunit of PI3K (phosphoinositide 3-kinase) [45,49]. Clearly, further analyses, e.g. using in vivo models, are required to determine the biological role of VEGFR1-interacting signal transducers. The exact role for VEGFR1 in endothelial cells, apart from serving as a reservoir for VEGF, is disputed. Several studies imply that VEGFR1 is dispensable for proliferation or migration of endothelial cells in vitro [50–52]. On the other hand, VEGFR1neutralizing antibodies have been employed to implicate VEGFR1 in endothelial cell actin reorganization and migration [53] through RACK1 (receptor for activated C-kinase 1) [54]. Endothelial cell differentiation and organization into vascular tubes may involve VEGFR1-dependent activation of PI3K/Akt [55].

171

VEGFR1 may regulate endothelial cells via cross-talk with VEGFR2, through dimerization, through transphosphorylation independently of heterodimerization and through regulation of receptor expression levels. Thus VEGFR1 and VEGFR2 form heterodimers on co-expressing cells [56], at least in vitro. VEGFR1–VEGFR2 heterodimers are predicted by computational modelling to comprise 10–50 % of the active, signalling, VEGF receptor complexes, and to form at the expense of VEGFR1 homodimers when VEGFR2 populations are larger [57]. VEGFR1–VEGFR2 heterodimers would form as a consequence of VEGFA binding, but not PlGF or VEGFB binding, since these latter ligands bind only to VEGFR1. Still, VEGFR1 can transphosphorylate VEGFR2 in response to PlGF, leading to sensitization of VEGFR2 to subsequent activation by VEGFA [46]. Furthermore, VEGFR1 may modulate VEGFinduced mitogenic signalling via VEGFR2 at least in part through altered VEGFR2 expression levels [51,58]. In haemangiomaderived endothelial cells, low expression of VEGFR1 is associated with constitutive VEGFR2 signalling [59]. In monocytes, VEGFR1-specific ligands VEGFB and PlGF induce signalling pathways known to operate downstream of most tyrosine kinase receptors such as ERK (extracellular-signalregulated kinase)/MAPK (mitogen-activated protein kinase), PI3K/Akt and the stress kinase p38MAPK. These signal transducers, as well as NFAT1 (nuclear factor of activated T-cells 1), have been implicated in the regulation of monocyte chemotaxis [60,61] (Figure 2). Vascular smooth muscle cells may respond to PlGF via VEGFR1, in particular under hypoxia, resulting in activation of the ERK1/2 and the JAK (Janus kinase)/STAT (signal transducer and activator of transcription) pathways [62]. VEGFR1 may also transduce signals for migration and invasion of several cancer cell lines, via the cytoplasmic tyrosine kinase Src and ERK1/2 pathways [63,64].

VEGFR1 biology: in vivo models and disease

Mouse models expressing recombinant variants of VEGFR1 have been instrumental in revealing its complex biology. Flt1 − / − (vegfr1 − / − ) embryos die at E9.0 (embryonic day 9) due to increased proliferation of endothelial cells and a severe disorganization and dysfunction of the vascular system [65]. Deletion of the intracellular domain including the VEGFR1 TK (tyrosine kinase) domain is, however, compatible with vascular development [66]. This suggests that in endothelial cells, VEGFR1 as well as sVEGFR1 may act as a VEGF decoy, serving to spatially control VEGFR2 signalling and formation of angiogenic sprouts [67]. Importantly, mice expressing a TK-deficient VEGFR1 (VEGFR1 TK − / − ) fail to mount an inflammatory response in pathologies such as cancer (see below). Membrane insertion of VEGFR1 appears to be important, as 50 % of mice expressing only soluble VEGFR1 as a result of deletion of both the transmembrane and the intracellular domains die in utero [68]. The negative regulatory role of VEGFR1 during vascular development may be exerted by sVEGFR1 (sFlt1), acting by sequestering VEGFA from signalling receptors or by the formation of non-functional heterodimers with VEGFR2 [69]. sFlt1 has important physiological functions such as in vascular maturation [70] and in the maintenance of corneal avascularity [71]. sFlt1 is expressed in the placenta in a regulated manner during gestation [72]. Overexpression of sFlt1 has been implicated in the aetiology of pregnancy-induced hypertension  c The Authors Journal compilation  c 2011 Biochemical Society

172

S. Koch and others

(pre-eclampsia) [73]. Plasma sFlt1 levels are also elevated in other diseases, such as cancer and ischaemia [74,75]. In addition to its negative regulatory role in vascular development, VEGFR1 is important in mounting an inflammatory response and inflammation-associated angiogenesis (denoted ‘pathological angiogenesis’) through recruitment of bonemarrow-derived myelomonocytic cells followed by deposition of angiogenic growth factors. Consequently, VEGFR1 TK − / − mice show impaired angiogenesis in a wide range of pathological conditions such as tumour growth [76], metastatic spread of tumours [77], choroidal neovascularization [78], rheumatoid arthritis [79] and cerebral ischaemia [80]. Regulation of inflammatory cell recruitment via VEGFR1 appears to be exerted mainly through PlGF. Thus PlGF is dispensable for embryonic and adult physiological angiogenesis [81], but it is crucial for inflammation-associated angiogenesis in a number of diseases (see [82] for a more detailed review on the properties of PlGF). VEGFB, another VEGFR1-specific ligand, plays a restricted role in angiogenesis in vivo [83]. Instead, VEGFB serves to promote fatty-acid uptake in endothelial cells, which is important in organs with high metabolic stress, such as the heart [84]. PlGF does not compensate for the loss of VEGFB in fattyacid transport, neither in vitro [84] nor in vivo as evidenced by the different phenotypes of vegfb- and plgf -gene-targeted mice [81,85,86]. VEGFB-deficient animals survive development, but display reduced heart size and impaired recovery after cardiac ischaemia [85,86]. The molecular mechanisms underlying these different functions of the two VEGFR1-specific ligands PlGF and VEGFB (i.e. co-receptors, restricted expression, etc.) remain to be identified. Internalization of VEGFR1

Activated VEGFR1 is most probably internalized through clathrin-mediated endocytosis as it forms a ternary complex with CBL (Cas-Br-M murine ecotropic retroviral transforming sequence homologue) and the adaptor protein CD2AP (CD2associated protein); the complex in turn associates with clathrin and the early endosome GTPase Rab (Ras-related GTPase) 4 [87]. VEGFR1 does not appear to become recycled, in contrast with VEGFR2 (see below), which is compatible with an important role for VEGFR1 in negative regulation through rapid clearance of VEGFA. VEGFR2 SIGNALLING Structure and expression of VEGFR2

VEGFR2 [also known as KDR in the human and Flk1 (fetal liver kinase-1) in the mouse] is a 210–230 kDa glycoprotein that binds VEGFA with a 10-fold lower affinity than VEGFR1 [40,88]. In addition to VEGFA, VEGFR2 binds proteolytically processed VEGFC and VEGFD [89]. Ligand binding involves extracellular Ig-like domains 2 and 3 of VEGFR2 (Figure 1). The parapox virus open-reading-frame-encoded VEGFE is the only known ligand to uniquely bind to VEGFR2. Alternative splicing results in the generation of sVEGFR2, which is present in various tissues such as the skin, heart, spleen, kidney and ovary, and in plasma. sVEGFR2 binds VEGFC and prevents binding to VEGFR3, consequently inhibiting lymphatic endothelial cell proliferation [90]. Progression of neuroblastoma is accompanied by down-regulation of sVEGFR2 [90–92]. sVEGFR2 may also contribute to vessel maturation by regulating mural cell migration and vessel coverage [70].  c The Authors Journal compilation  c 2011 Biochemical Society

VEGFR2 is expressed most prominently in vascular endothelial cells and their embryonic precursors, with highest expression levels during embryonic vasculogenesis and angiogenesis [93– 95]. VEGFR2 is also found in a range of non-endothelial cells such as pancreatic duct cells, retinal progenitor cells, megakaryocytes and haemopoietic cells (see [95] and references therein). VEGFR2 expression is induced in conjunction with active angiogenesis, such as in the uterus during the reproductive cycle [97], and in pathological processes associated with neovascularization, such as cancer [98,99]. VEGFR2 expression on tumour cells has been noted for melanoma and haematological malignancies [100]. The detailed expression and functionality of VEGFR2 in different cell types in the adult remains to be determined. Interestingly, Maharaj et al. [101] detected constitutively phosphorylated VEGFR2 in vivo in adult mouse tissues including the liver, lung, adipose tissue and kidney.

VEGFR2 signal transduction

VEGFR2 is known to transduce the full range of VEGF responses in endothelial cells, i.e. regulating endothelial survival, proliferation, migration and formation of the vascular tube. Major phosphorylation sites in VEGFR2 are Tyr951 in the kinase insert domain, Tyr1054 and Tyr1059 within the kinase domain, and Tyr1175 and Tyr1214 in the C-terminal domain [102,103] (Figure 3). Additional phosphorylation sites in VEGFR2 have been identified at positions 1223, 1305, 1309 and 1319 [102], but their function(s) are as yet unclear. Moreover, the juxtamembrane tyrosine residue Tyr801 is phosphorylated in isolated intracellular domains [104]; whether Tyr801 is phosphorylated in the intact VEGFR2 remains to be shown. Phosphorylated Tyr951 serves as a binding site for the SH2-domain-containing TSAd (T-cell-specific adapter molecule) [102,105], also denoted VRAP (VEGFR-associated protein; however, TSAd is the consensus designation). Tyr951 is phosphorylated during active angiogenesis [102], but not in quiescent endothelial cells. Mutation of VEGFR2 at Tyr951 as well as TSAd knockdown prevent VEGFA-dependent actin reorganization and migration, but not proliferation, of endothelial cells. TSAd forms a complex with Src in a VEGFA-dependent manner [102]. Phosphorylation at Tyr1054 and Tyr1059 , which are located in the kinase domain activation loop, are critical for kinase activity [106]. Phosphorylation of these residues may be preceded by autophosphorylation at Tyr801 [104]. The positive regulatory tyrosine residues in receptor tyrosine kinases do not as a rule participate in binding of SH2-domain proteins. However, Tyr1059 has been shown to bind Src, which in turn phosphorylates other residues in VEGFR2, such as Tyr1175 , as well as downstream signal transducers, such as the actin-binding protein IQGAP1 (IQ-motifcontaining GTPase-activating protein 1) [107]. IQGAP1 has been implicated in the regulation of cell–cell contacts, proliferation and migration [108]. Phosphorylation of Tyr1175 creates a binding site for PLCγ [44,103], but it also binds the adapter proteins SHB (SH2domain-containing adaptor protein B) and Sck [SHC (Src homology and collagen homology)-related adaptor protein] [109,110]. SHB binds to FAK (focal adhesion kinase), which is tyrosine phosphorylated in VEGF-treated cells [111] and contributes to endothelial cell attachment and migration [112]. Tyr1175 has also been described to bind SHC or GRB2, which recruit the nucleotide-exchange factor SOS (Son of sevenless) to VEGFR2 [110,113]. This pathway may control Ras activation and mitogenicity in response to VEGF [114]. VEGF-induced Ras

Signal transduction by VEGF receptors

Figure 3

173

VEGFR2 tyrosine phosphorylation sites and signal transduction

Schematic outline of activated and dimerized VEGFR2. Upon VEGFA binding to extracellular Ig-like domains 2 and 3 of VEGFR2 (hatched circles), signalling molecules (‘rocket’ shapes) bind to respective tyrosine phosphorylation sites (indicated by numbers) in the intracellular domain of VEGFR2 and activate downstream mediators (ovals). Tyr1054 and Tyr1059 are crucial for VEGFR2 kinase activity (italics). The complex network of intracellular signal transduction pathways results in biological responses such as proliferation, migration, survival and permeability (bottom boxes), which are all required for the co-ordinated arrangement of endothelial cells in three dimensions to form and maintain vascular tubes. See the main text for details. CDC42, cell division cycle 42; DAG, diacylglycerol; IP3 , inositol 1,4,5-trisphosphate; MEK, MAPK/ERK kinase; PIP2 , phosphatidylinositol 4,5-bisphosphate; SOS, Son of sevenless.

activation has been implicated in the generation of pleiotropic lipid mediators, the prostaglandins [e.g. PGI2 (prostaglandin I2 )] via cPLA2 (cytoplasmic phospholipase A2 ) (see [115] and references therein). However, an important pathway for VEGFregulated proliferation appears to be through Tyr1175 -PLCγ mediated activation of PKC (protein kinase C) and downstream induction of the ERK pathway leading to cell proliferation [103]. The PKC isoforms PKCβ [116,117], PKCα and PKCζ have been shown to mediate VEGF-induced proliferation [118]. PKCs may also mediate activation of SPK (sphingosine kinase) [119]. Activation of PKCs downstream of VEGF results in serine phosphorylation and activation of PLCβ3, an effector of heterotrimeric G-proteins [120]. PKC activation also results in activation of PKD (protein kinase D) [121], which promotes nuclear translocation of HDAC (histone deacetylase) 5 and HDAC7 [122,123], followed by phosphorylation of CREB (cAMP-response-element-binding protein) as well as HSP (heatshock protein) 27 [124]. Tyr1214 allows recruitment of Nck and the cytoplasmic tyrosine kinase Fyn to VEGFR2. A Nck–Fyn complex mediates phosphorylation of PAK (p21-activated protein kinase) 2 and subsequent activation of Cdc42 (cell division cycle 42) and p38MAPK [125]. HSP27 is also implicated downstream of Tyr1214 , in a MAPKAPK2 (MAPK-activated protein kinase 2)-dependent manner [126]. It has not yet been fully elucidated which phosphorylation sites or domains in VEGFR2 are important for generation of lipid mediators. The docking protein GAB (GRB2-binding protein) 1 contains a binding site for the p85 subunit of PI3K. VEGF-

induced recruitment of GAB1 leads to PIP3 (phosphatidylinositol 3,4,5-trisphosphate) generation, which in turn activates the small GTPase Rac, resulting in membrane ruffle formation and cell motility [127]. Flow shear stress has been shown to stimulate VEGFR2 activation and GAB1 tyrosine-residue phosphorylation in a VEGF-independent manner [128]. PI3K in turn promotes activation of PKB (protein kinase B, also called Akt) through PDK (phosphoinositide-dependent kinase) 1 and PDK2 (Figure 3). Akt phosphorylates BAD [Bcl (Bcell lymphoma)-2-associated death promoter] and caspase 9 and thereby inhibits their apoptotic activity [129,130]. VEGFA also induces the expression of the anti-apoptotic proteins Bcl-2 and A1 [131,132], as well as the IAP (inhibitors of apoptosis) family members XIAP (X-chromosome-linked IAP) and survivin, which inhibit the terminal effector caspases 3 and 7 [133,134].

VEGFR2 biology: in vivo models and disease

Vegfr2 − / − mice die at E8.5 from impaired development of haemopoietic and endothelial cells [135], a phenotype similar to that of the vegfa − / − mouse [136,137]. VEGFR2 is therefore believed to be the main transducer of VEGFA effects on endothelial cell differentiation, proliferation, migration and formation of the vascular tube. The Tyr1173 phosphorylation site (corresponding to Tyr1175 in humans) is crucial for endothelial and haemopoietic cell development. Mice with a Y1173F knockin die between E8.5 and E9.5 from lack of endothelial and haemopoietic  c The Authors Journal compilation  c 2011 Biochemical Society

174

S. Koch and others

development [138], similar to the phenotype of vegfr2 − / − mice [135]. The phenotype of PLCγ 1-deficient mice mirrors that of Y1173F mice [139,140]. Moreover, PLCγ 1 is required for VEGF responsiveness and arterial development in the zebrafish [141]. The loss of binding of the adaptors SHB and Sck may also contribute to the vegfr2 y1173f-like phenotype. SHB is not required for vascular development, but SHB-deficient mice show defects in vessel functionality and impaired tumour growth [142]. In contrast, mice with a VEGFR2 Y1212F substitution (corresponding to position 1214 in the human) [138] or a Y949F substitution (corresponding to Tyr951 in the human; X. Li and L. Claesson-Welsh, unpublished work) are viable and fertile. Regulation of vascular permeability may be one of the main in vivo functions of VEGFA, in keeping with its original designation, VPF (vascular permeability factor). Vascular permeability is essential for normal tissue homoeostasis, and dysregulated permeability leading to chronic oedema is an important feature in cancer and in circulatory diseases such as peripheral ischaemia and heart failure. Fenestrae, which can be induced by VEGFA, are found in small numbers in many types of vascular endothelium and are especially numerous in specialized vascular beds in hormone-secreting tissues [143]. VEGFA-induced extravasation of proteins or white blood cells is mediated by VEGFR2 in vivo [144]. The acute increase in vascular permeability induced by VEGFA occurs through venules, not capillaries [145]. Two major mechanisms have been implicated in vascular permeability; creation of transcellular endothelial pores and transient opening of paracellular endothelial junctions [146]. Transcellular pores such as the VVO (vesiculo-vacuolar organelle) are believed to be assembled by lipid microdomains denoted caveolae that form vesicles/vacuoles spanning the venular endothelium [147]. VEGFR2 is localized in caveolae and occurs in complex with caveolin-1 through its C-terminal tail [148]. Caveolin-1-targeted mice, which lack caveolae, retain VVOs and display increased permeability to plasma albumin and increased tumour growth [149]. Transient opening of endothelial cell–cell junctions in response to VEGF is well documented in vitro and involves dissolution of tight junctions as well as adherens junctions [150]. VEGF dissolves adherens junction complexes, composed of VE-cadherin (vascular endothelial cadherin), β-catenin, p120-catenin, and αcatenin [151] through activation of the cytoplasmic tyrosine kinases Src and Yes [152]. Src and Yes phosphorylate VEcadherin [153,154] and β-catenin [155,156], leading to increased permeability in cell culture models [152,157] and in mice [154]. Enhanced tyrosine phosphorylation of adherens junction proteins may also be a consequence of reduced protein tyrosine phosphatase activity [158,159]. VEGF may furthermore regulate junctional permeability by stimulating endocytosis of VEcadherin. This pathway is initiated through activation of Src, which via PAK2 phosphorylates VE-cadherin on serine residues [160]. VEGF-induced permeability also involves eNOS (endothelial nitric oxide synthase)-mediated generation of NO [161]. eNOS is activated either by PLCγ -dependent Ca2 + influx or Akt-mediated phosphorylation at Ser1179 of eNOS [162,163]. Impaired VEGFinduced permeability in eNOS-deficient mice supports a critical role for eNOS-derived NO in the regulation of VEGF-induced vascular permeability [161]. NO in turn regulates microvascular fluid and protein exchange at postcapillary venules by modulating blood flow [164] or by acting, in a cGMP-PKG (protein kinase G)dependent manner, on the cytoskeleton to increase the passage of macromolecules [165]. Furthermore, NO-induced S-nitrosylation of β-catenin at Cys619 promotes VEGF-induced endothelial cell permeability [166].  c The Authors Journal compilation  c 2011 Biochemical Society

Internalization of VEGFR2

Activation of VEGFR2 leads to dissociation from caveolin-1 [148,167] and translocation to endosomes [168]. Alternatively, VEGFR2 is translocated to perinuclear caveosomes through caveolar endocytosis [148,169,170]. The degree of internalization of VEGFR2 in vitro is dependent on cell–cell junctions [171] and may be regulated by other receptor tyrosine kinases such as ephrin-B2 in an as yet unclear manner [172]. The NRP-binding adaptor synectin and its interaction with myosin VI are required to move VEGFR2-containing endosomes away from the plasma membrane, where PTP1b dephosphorylates VEGFR2 specifically at the Tyr1175 site [173]. Internalized VEGFR2 may be degraded or recycled to the plasma membrane. VEGFR2 is routed for degradation as a consequence of post-translational modifications, i.e. CBL-mediated ubiquitination [171,174,175] or PKC-mediated serine-residue phosphorylation of VEGFR2 at Ser1188 and Ser1191 in the Cterminal tail [176]. Alternatively, activated abluminal VEGFR2 may exit caveolae and associate with ARF (ADP-ribosylation factor) 6 and RAC1 (Ras-related C3 botulinum toxin substrate 1) in focal adhesions [177]. ARF6 is known to promote plasma membrane recycling and co-ordinate actin remodelling [178]. VEGFR2 signalling is not terminated in conjunction with internalization; instead, the receptor continues to signal from endosomes and during sorting until reaching its final destinations [148,179]. Indeed, internalization of VEGFR2 is required for ERK and Akt activation, whereas p38MAPK is activated only by surface VEGFR2 [7,171,172]. Unlike other receptor tyrosine kinases, VEGFR2 is stored in intracellular Rab4- and Rab11-negative vesicles in resting endothelial cells from where it is delivered to the plasma membrane in a Src-dependent manner, in response to VEGFA [175].

VEGFR3 SIGNALLING Structure and expression of VEGFR3

VEGFR3 (also denoted Flt4) is synthesized as a precursor protein of 195 kDa. The precursor is proteolytically cleaved in the fifth Ig-like domain, generating an N-terminal peptide, which remains disulfide-bonded to the mother protein [180] (Figure 1). VEGFR3 is activated by the binding of VEGFC and VEGFD. Proteolytic processing of VEGFC and VEGFD results in increased affinity for binding to both VEGFR2 and VEGFR3 [181,182]. Processed VEGFC binds with higher affinity to VEGFR3 than to VEGFR2 (for hVEGFC–hVEGFR3, K d = 0.44 × 10 − 8 M, and for hVEGFC–hVEGFR2, K d = 2.2 × 10 − 8 M, where h is human) [183]. VEGFC binding requires Ig-loops 1 and 2 in VEGFR3 [184], whereas binding to VEGFR2 involves loops 2 and 3 [185]. hVEGFD shows similar affinity for both VEGFR2 and VEGFR3; in contrast, mouse VEGFD binds only to VEGFR3 [186]. This is the only documented example of species-specific patterns of VEGF–receptor interactions. Furthermore, N-terminal residues in VEGFD are essential for VEGFR3, but not VEGFR2, activation [187]. Cloning of VEGFR3 identified alternative C-terminal splicing, giving rise to short and long VEGFR3 isoforms [188,189]. The short VEGFR3 isoform lacks the 65 C-terminal amino acid residues due to a retroviral insertion and occurs only in the human genome. The long isoform is predominantly expressed, but, interestingly, in thyroid diseases, the short variant is abundantly expressed. This may suggest a different function of the two isoforms [190]. Notably, the short isoform lacks the two most C-terminal phosphorylation sites, which are phosphorylated by

Signal transduction by VEGF receptors

Figure 4

VEGFR3 tyrosine phosphorylation sites and signal transduction

Schematic outline of activated and dimerized VEGFR3. Tyr1068 (italics) is crucial for VEGFR3 kinase activity. VEGFR3 occurs as a short and a long isoform (the length varying in the region indicated by *).Tyr1337 and Tyr1363 are transphosphorylated only in the long isoforms by VEGFR3 in homodimers, but not by VEGFR2 in heterodimers. VEGFR3 may contribute to proliferation, migration and survival of lymphendothelial cells through ligand-dependent or ligand-independent mechanisms. See the main text for details. ECM, extracellular matrix.

VEGFR3 in receptor homodimers, but not by VEGFR2 in VEGFR2–VEGFR3 heterodimers [191]. VEGFR3 expression starts in the primary vascular plexus at E8.5 during mouse development [192]. At a later stage of embryogenesis, VEGFR3 is expressed in venous endothelial cells in the cardinal vein, which subsequently gives rise to VEGFR3expressing lymphatics [193]. VEGFR3 has an essential role in lymphatic endothelial cells, but its expression is induced in endothelial cells in conjunction with active angiogenesis [194], such as in the tumour vasculature or in endothelial tip cells of angiogenic sprouts in the developing retina [195]. Moreover, zebrafish VEGFR3 is expressed in the tip cells of developing intersegmental arterial vessels [196]. VEGFR3 is also expressed in non-endothelial cells such as osteoblasts [197], neuronal progenitors [198] and macrophages [199]. Whether VEGFR3 is expressed in tumour cells is disputed [200]. VEGFR3 signal transduction

Binding of VEGFC/VEGFD to VEGFR3 leads to kinase activation and phosphorylation of at least five C-terminal tyrosine residues in VEGFR3 (Tyr1230 , Tyr1231 , Tyr1265 , Tyr1337 and Tyr1363 ) [191] (Figure 4). Mutational analyses show that residues Tyr1063 and Tyr1068 , located in the VEGFR3 kinase domain activation loop, are fundamental for kinase activity [201]. The roles of the different phosphorylation sites have been studied in vitro. As mentioned above, tyrosine residues in the kinase activation loop are not considered as accessible for SH2protein binding. Still, phosphorylated Tyr1063 has been shown to interact with the adaptor protein CRK I/II (C10 regulator of kinase; originally identified in a C10 retrovirus), which activates the JNK (c-Jun N-terminal kinase) pathway leading to cell

175

survival [201]. On the other hand, mutation of Tyr1337 leads to reduced binding of SHC–GRB2 and reduced mitogenic signalling [202]. Phosphorylation of Tyr1230 and Tyr1231 creates a docking site for SHC–GBR2, which promotes signalling in the ERK1/2 and PI3K/Akt pathways [201]. The PI3K pathway is critical in the development of lymphatics by regulating lymphendothelial migration [183]. Migration of lymphendothelial cells in vitro is induced by both VEGFC and a mutant VEGFC156S that binds specifically to VEGFR3 and not to VEGFR2. Integrins can induce SRC-dependent phosphorylation of VEGFR3 at tyrosine residues 830, 833, 853, 1063, 1333 and 1337. This phosphorylation is independent of VEGFC/VEGFD and activation of the VEGFR3 kinase. Integrin-mediated VEGFR3 phosphorylation leads to recruitment of the adaptor proteins CRKI/II and SHC to VEGFR3, thereby inducing activation of JNK [203]. These findings suggest that blocking receptor ligands is not sufficient to suppress VEGFR3 signalling. There are several studies describing heterodimers between VEGFR2 and VEGFR3 [16,191,204] (Figure 1) and, interestingly, the heterodimers appear to be functionally different from homodimers of each receptor. Indeed, it has been stated that VEGFR3 needs to be associated with VEGFR2 to induce at least certain VEGFC and VEGFD-dependent cellular responses [204]. On the other hand, VEGFR3 homodimers have been implicated in three-dimensional organization of endothelial cells and lumen formation [205]. VEGFR2–VEGFR3 heterodimers appear to be biologically active in angiogenesis, on the basis of treatment with receptor-neutralizing antibodies [16]. That VEGFR2– VEGFR3 heterodimers are functionally distinct is indicated by the fact that heterodimeric VEGFR3 is phosphorylated only at three of its five potential C-terminal tyrosine phosphorylation sites; the two most C-terminal tyrosine residues appear not to be accessible for the VEGFR2 kinase [191].

VEGFR3 biology: in vivo models and disease

VEGFC and VEGFR3 are critical regulators of lymphendothelial function. Patients with certain variants of hereditary lymphoedema carry mutations in VEGFR3, resulting in an inactive tyrosine kinase [206–208]. In addition, the Chy mouse mutant, whose phenotype is characterized by accumulation of chylous ascites, carries a point mutation in the VEGFR3 tyrosine kinase domain, which renders the kinase enzymatically inactive [209]. The phenotype of naturally occurring VEGFR3 mutants strongly indicates that, physiologically, VEGFR3 signal transduction is critical in the regulation of lymphatic vessel function. During development, however, VEGFR3 plays an important role in blood vascular development. Vegfr3 − / − mice die at E10.5, before the formation of lymphatics, due to cardiovascular failure and impaired hierarchical organization of the vasculature [210]. Gene inactivation to eliminate expression of VEGFC alone, or both VEGFC and VEGFD, unexpectedly results in defects mainly in lymphatic vessels, whereas blood vessels are unaffected [211]. Studies on recombinant mice expressing mutated or truncated forms of VEGFR3 have shown that VEGFR3 has to be expressed and functional to allow lymphatic development. In contrast, blood vessel development is maintained in mice expressing a kinase-deficient VEGFR3 retaining the extracellular ligand-binding region [212]. These results support the concept that VEGFR3 expression in blood vascular endothelial cells regulates VEGFR2 signalling by heterodimerization. VEGFR3 may also regulate VEGFR2 signalling by trapping VEGFC.  c The Authors Journal compilation  c 2011 Biochemical Society

176

S. Koch and others

The importance of the VEGFR2–VEGFR3 cross-talk is furthermore underscored by the effects of VEGFR3 neutralizing antibodies, which inhibit both VEGFR3 homodimerization and heterodimerization with VEGFR2, resulting in suppression both of vascular and lymphatic endothelial cell migration and vessel sprouting [213]. In vivo, VEGFC-induced migration and sprouting of lymphendothelial precursor cells from restricted regions of the cardinal vein is required for development of the lymphatic system [214] and, as discussed above, PI3K may be critical for lymphendothelial migration [183]. An important role for the PI3K pathway downstream of VEGFR3 is supported by in vivo results. The gene Pik3r1 encodes three proteins (p85α, p55α and p50α) that serve as regulatory subunits of class IA PI3Ks. Gene inactivation leading to loss of all protein products of this gene results in chylous ascites, indicative of impaired lymphatic function [215]. Furthermore, VEGFC-mediated Akt activation is required for embryonic and adult lymphangiogenesis [216]. Finally, Ras proteins regulate VEGFR3 expression and activity in lymphatic signalling [217].

Internalization of VEGFR3

Similar to VEGFR2, VEGFR3 internalization is crucial for proangiogenic downstream signalling. VEGFR3 internalization and signalling involving RAC1, Akt and ERK is regulated by ephrin B2 [218].

VEGF CO-RECEPTORS

VEGF co-receptors are broadly defined in the present review as VEGF-binding cell-surface-expressed molecules that are devoid of intrinsic catalytic activity, but which modulate the signal-transduction output downstream of VEGFRs. However, we do not categorically exclude that signals may be transmitted downstream of co-receptors also independently of VEGFRs.

HS

HS and the more abundantly sulfated, mast-cell-derived, H (heparin) modulate VEGF biology in several ways by binding not only VEGF, but also receptors and co-receptors such as NRP1 (Figure 5A). All VEGFA isoforms except VEGFA121 and the VEGFAxxxb forms bind HS/H [219]. Moreover, the fourth Igloop of VEGFR1 and the sixth and seventh Ig-loops of VEGFR2 have been shown to directly interact with HS/H [220,221]. NRP1, but not NRP2 (see below), binds HS/H [222]. Simultaneous HS/Hbinding to several or all components in the signalling complex is expected to increase its stability [223]. Moreover, presentation of VEGFA165 to VEGFR2 in trans, by HSPGs (HS proteoglycans) expressed on adjacent cells such as pericytes, further increases the signalling amplitude and duration [224] (Figure 5B), most probably by blocking internalization of the receptor. HS serves as a reservoir for growth factors, and controlled release allows formation of growth-factor gradients [225]. Thus tip cells of sprouting blood vessels migrate in response to VEGFA164 gradients, and these gradients are shaped by interactions with HSPG [226]. The role of HS in regulating retinal vascular development has been studied in mice that selectively express single VEGF isoforms. Interestingly, VEGFA164/164 mice (expressing only VEGFA164 and none of the other VEGFA isoforms) show normal vascular development, whereas VEGFA120/120 mice (expressing only VEGFA120 and  c The Authors Journal compilation  c 2011 Biochemical Society

none of the other VEGFA isoforms) exhibit severe defects in vascular outgrowth and patterning [227]. NRPs

There are two NRP homologues, NRP1 and NRP2, which are 130 kDa transmembrane proteins with a small cytoplasmic tail lacking intrinsic catalytic function [228]. NRP1 may homomultimerize and can also form heteromultimers with NRP2 [229]. There are several alternatively spliced NRP1 and NRP2 isoforms [230]. NRPs were first identified as receptors for class 3 semaphorins, a family of soluble molecules with neuronal guidance functions, implicated in modulating the development of the nervous and vascular systems [231–233]. NRP1 was later shown to bind exon 8-encoded regions of VEGFA isoforms such as VEGFA165. Heparin-mediated VEGFR2–NRP1 complex formation depends on heparin-binding regions both in VEGFA isoforms (encoded by exons 6 and 7) and in NRP1 (b1b2) [234]. Therefore VEGFA121 (lacking exon 7) also binds directly to NRP1, but without promoting complex formation between NRP1 and VEGFR2 [235]. NRP2 on the other hand associates with VEGFR3 in a VEGFC- or VEGFD-dependent manner [236]. The NRPs are organized in extracellular subdomains (Figure 5): the a1-a2 domain is homologous with components of the complement system and binds semaphorins; the b1-b2 domain shares homology with coagulation factors V and VIII, binds VEGFA and HS/H, and supports semaphorin-binding to the a1-a2 domain; and the c domain is homologous with the MAM (protease maprin, A5-protein, PTPμ) domain [230]. Both NRP homologues bind VEGFA165, but with different affinities [237]. In addition, NPR1 binds VEGFB, PlGF and VEGFE, whereas NRP2 binds VEGFA145 and VEGFC [230]. NRP1 is modified, in a cell-type-specific manner, by heparan and chondroitin sulfation, which may allow indirect binding of growth factors other than VEGFA to the glycosaminoglycan side chains [238]. NRP1 modulates VEGFR signalling, leading to enhanced migration [234] and survival [239,240] of endothelial cells in vitro. Furthermore, NRP1 has been implicated in VEGFR2-mediated permeability [241] and in VEGFA-induced three-dimensional biology, such as vessel sprouting and branching [242], NRPs have also been strongly implicated in intracellular trafficking of VEGFR2 by association of the NRP1 C-terminal SEA (Ser-Glu-Ala, the last three residues in the NRP1 C-terminal tail) motif to the PDZ domain of the adaptor synectin (also called GIPC) [240,243] (Figure 5A). The exact molecular mechanism whereby NRP1 modulates VEGF biology is unclear, but altered VEGFR2 signalling in the presence of NRP1 of both of p38MAPK [242] and p130CAS (Crk-associated substrate) [244] have been reported. It remains to be shown whether NRP1 transduces VEGF signals independently of VEGFR2 through its C-terminal PDZbinding domain or whether it modulates VEGFR2 signalling, e.g. by influencing the tertiary structure of the receptor or the stability of the signalling complex. During development, NRP1 is preferentially expressed in arteries, and NRP2 in veins and lymphatics [245–247]. Both overexpression and disruption of NRP1 in mice lead to embryonic lethality at E12.5–13.5 due to either an excess of vessel formation or a range of vascular abnormalities respectively [248]. NRP2knockout mice show normal development of arteries and veins, but exhibit marked deviations in the development of capillaries and small lymphatics [247]. In agreement, NRP2 mediates VEGFC-induced lymphatic sprouting in a VEGFR3-dependent manner [249]. NRP1/NRP2-double-knockout mice display a more severe phenotype and die in utero at E8.5 with a completely

Signal transduction by VEGF receptors

Figure 5

177

Schematic outline of the interactions of VEGFR2 with its co-receptors HS/H, NRP1 and integrins

(A) VEGFA (green) bridges VEGFR2 (red) and NRP1 (yellow) with its exon 4- or exon 8-encoded regions respectively. The exon 6-encoded domain of VEGFA interacts with HS/H (brown) and exons 1–5-encoded regions are involved in binding VEGFR2. VEGFA-binding domains of VEGFR2 are indicated by vertically hatched circles. HS/H bind VEGFA, VEGFR2 and NRP1. HS/H-binding domains in VEGFR2 and NRP1 are indicated by horizontally hatched circles. The VEGFA- and HS/H-binding domains in NRP1 (b1 and b2) overlap. Proteoglycans (brown) can be soluble or plasma membrane anchored. The intracellular SEA motif of NRP1 binds the PDZ domain of synectin, which mediates internalization of the signalling complex via myosin VI. (B) Alternatively, VEGFA can be presented to VEGFR2 in trans through HSPGs located on adjacent cells, possibly leading to altered signalling and arrest of the signalling complex in the plasma membrane. (C) Integrin αVβ3 (purple) can bind VEGFR2 [the exact domain(s) are unknown] in a VEGFA-dependent manner and support the attachment of the cell to the extracellular matrix (ECM) through vitronectin. Matrix-bound VEGFA prolongs VEGFR2 phosphorylation at Tyr1214 and downstream p38MAPK activation dependent on β1 integrin (pink) association. Downstream mediators include FAK and vinculin, which initiate cell migration. Integrin αVβ3 may also sequester NRP1 to prevent interaction with VEGFR2. See the main text for details.

avascular yolk sac [250]. Treatment with a neutralizing anti-NRP1 antibody reduces angiogenesis and vascular remodelling in vivo and has an additive effect with anti-VEGFA antibody therapy in reducing tumour growth and tumour vessel organization [251]. Treatment with a neutralizing anti-NRP2 antibody reduces lymphangiogenesis and functional lymphatics in tumours, while leaving established lymphatics unaffected. The reduced lymphatic function is accompanied by reduced metastatic spread [252]. Soluble forms that may serve to sequester VEGF have been identified for both neuropilin homologues [230,253]. In agreement, mice overexpressing soluble NRP1 show decreased vascular permeability [230].

Integrins

Integrins are transmembrane heterodimers that mediate cellmatrix adhesion by specific binding to extracellular matrix components, such as collagen, fibronectin, vitronectin and laminin. In cells plated on vitronectin, VEGF induces complex formation between VEGFR2 and the αVβ3 integrin (Figure 5C). The β3 integrin subunit binds VEGFR2 through its extracellular domain and the complex is stabilized by the αV subunit [254,255]. The VEGFR2–αVβ3-integrin association is important for full VEGFR2 activity, for activation of p38MAPK and FAK, and for recruitment of actin-binding vinculin to initiate endothelial cell migration [256,257]. VEGFA-dependent binding of integrin αVβ3 to NRP1 may prevent interaction with VEGFR2 [258]. In vivo, the VEGFR2–αVβ3 interaction is required for tumourinduced and wound angiogenesis as well as in the recruitment of bone-marrow-derived cells to angiogenic sites [259,260]. Complex formation between VEGFR2 and β1 integrin induced by matrix-bound VEGF leads to prolonged phosphorylation of Tyr1214 of VEGFR2 and association of β1 integrin with focal adhesions [7].

Complex formation between VEGFR3 and αVβ1 integrins in response to VEGFC treatment of cells plated on fibronectin has also been reported [261].

CONCLUDING REMARKS

VEGFR signalling affects a number of vital processes during development, in adult physiology and pathology. Therefore studies on VEGFR signalling have received ample interest from a broad community of academia, the pharmaceutical industry and clinical medicine. However, the recent challenges in the application of VEGF-targeted cancer therapy, i.e. resistance to VEGF neutralization [262] and increased invasion and metastasis [263,264], have very clearly shown that we need to learn more about how VEGF affects blood vessels in intact tissues under different metabolic conditions (see [265] for a review on VEGFRtargeted therapy in cancer). Such information is urgently needed to control the impact of vascular targeting in diseases. VEGFR signalling has until now mostly been studied in vitro. These studies have provided important knowledge on the impact of VEGF on endothelial cell function, which in turn has promoted the development of anti-angiogenic drugs for treatment of diseases characterized by excess angiogenesis. However, a number of parameters of in vivo vascular function cannot be accurately modelled in monolayer cultures of immortalized, transformed or even primary endothelial cells. Such parameters include (i) that cell–cell junctions may not be properly organized, (ii) that an appropriate vascular basement membrane and supporting cells are missing, (iii) that the three-dimensional context may not be represented, and (iv) that co-receptors may not be adequately expressed. Therefore progress in research on VEGF receptor signalling now requires not only tools for sensitive and accurate biochemical analyses, but also ambitious in vivo models.  c The Authors Journal compilation  c 2011 Biochemical Society

178

S. Koch and others

FUNDING

This work was supported by Uppsala University, the Swedish Cancer Foundation [contract number 10 0543], the Swedish Science Council [contract number K2011-67X12552-14-5], the Knut and Alice Wallenberg Foundation, a postdoctoral grant from the Spanish Ministry of Science and Education (MEC) (to S.T.), a postdoctoral grant from the Spanish Association of Liver Diseases (AEEH) (to S.T.), and the Gustaf Adolf Johansson Foundation at Uppsala University (to S.K.). REFERENCES 1 Senger, D. R., Galli, S. J., Dvorak, A. M., Perruzzi, C. A., Harvey, V. S. and Dvorak, H. F. (1983) Tumor cells secrete a vascular permeability factor that promotes accumulation of ascites fluid. Science 219, 983–985 2 DiSalvo, J., Bayne, M. L., Conn, G., Kwok, P. W., Trivedi, P. G., Soderman, D. D., Palisi, T. M., Sullivan, K. A. and Thomas, K. A. (1995) Purification and characterization of a naturally occurring vascular endothelial growth factor·placenta growth factor heterodimer. J. Biol. Chem. 270, 7717–7723 3 Ferrara, N. (2010) Binding to the extracellular matrix and proteolytic processing: two key mechanisms regulating vascular endothelial growth factor action. Mol. Biol. Cell 21, 687–690 4 Lee, S., Jilani, S. M., Nikolova, G. V., Carpizo, D. and Iruela-Arispe, M. L. (2005) Processing of VEGF-A by matrix metalloproteinases regulates bioavailability and vascular patterning in tumors. J. Cell Biol. 169, 681–691 5 Grunewald, F. S., Prota, A. E., Giese, A. and Ballmer-Hofer, K. (2010) Structure–function analysis of VEGF receptor activation and the role of coreceptors in angiogenic signaling. Biochim. Biophys. Acta 1804, 567–580 6 Uniewicz, K. A. and Fernig, D. G. (2008) Neuropilins: a versatile partner of extracellular molecules that regulate development and disease. Front. Biosci. 13, 4339–4360 7 Chen, T. T., Luque, A., Lee, S., Anderson, S. M., Segura, T. and Iruela-Arispe, M. L. (2010) Anchorage of VEGF to the extracellular matrix conveys differential signaling responses to endothelial cells. J. Cell Biol. 188, 595–609 8 Harper, S. J. and Bates, D. O. (2008) VEGF-A splicing: the key to anti-angiogenic therapeutics? Nat. Rev. Cancer 8, 880–887 9 Ogawa, S., Oku, A., Sawano, A., Yamaguchi, S., Yazaki, Y. and Shibuya, M. (1998) A novel type of vascular endothelial growth factor, VEGF-E (NZ-7 VEGF), preferentially utilizes KDR/Flk-1 receptor and carries a potent mitotic activity without heparin-binding domain. J. Biol. Chem. 273, 31273–31282 10 Yamazaki, Y., Tokunaga, Y., Takani, K. and Morita, T. (2005) C-terminal heparin-binding peptide of snake venom VEGF specifically blocks VEGF-stimulated endothelial cell proliferation. Pathophysiol. Haemost. Thromb. 34, 197–199 11 Germain, S., Monnot, C., Muller, L. and Eichmann, A. (2010) Hypoxia-driven angiogenesis: role of tip cells and extracellular matrix scaffolding. Curr. Opin. Hematol. 17, 245–251 12 Randi, A. M., Sperone, A., Dryden, N. H. and Birdsey, G. M. (2009) Regulation of angiogenesis by ETS transcription factors. Biochem. Soc. Trans. 37, 1248–1253 13 Ushio-Fukai, M. and Nakamura, Y. (2008) Reactive oxygen species and angiogenesis: NADPH oxidase as target for cancer therapy. Cancer Lett. 266, 37–52 14 Lee, S., Chen, T. T., Barber, C. L., Jordan, M. C., Murdock, J., Desai, S., Ferrara, N., Nagy, A., Roos, K. P. and Iruela-Arispe, M. L. (2007) Autocrine VEGF signaling is required for vascular homeostasis. Cell 130, 691–703 15 Bussmann, J., Lawson, N., Zon, L. and Schulte-Merker, S. (2008) Zebrafish VEGF receptors: a guideline to nomenclature. PLoS Genet. 4, e1000064 16 Nilsson, I., Bahram, F., Li, X., Gualandi, L., Koch, S., Jarvius, M., Soderberg, O., Anisimov, A., Kholova, I., Pytowski, B. et al. (2010) VEGF receptor 2/-3 heterodimers detected in situ by proximity ligation on angiogenic sprouts. EMBO J. 29, 1377–1388 17 Dosch, D. D. and Ballmer-Hofer, K. (2010) Transmembrane domain-mediated orientation of receptor monomers in active VEGFR-2 dimers. FASEB J. 24, 32–38 18 Stuttfeld, E. and Ballmer-Hofer, K. (2009) Structure and function of VEGF receptors. IUBMB Life 61, 915–922 19 Ruch, C., Skiniotis, G., Steinmetz, M. O., Walz, T. and Ballmer-Hofer, K. (2007) Structure of a VEGF–VEGF receptor complex determined by electron microscopy. Nat. Struct. Mol. Biol. 14, 249–250 20 Yang, Y., Xie, P., Opatowsky, Y. and Schlessinger, J. (2010) Direct contacts between extracellular membrane-proximal domains are required for VEGF receptor activation and cell signaling. Proc. Natl. Acad. Sci. U.S.A. 107, 1906–1911 21 Lemmon, M. A. and Schlessinger, J. (2010) Cell signaling by receptor tyrosine kinases. Cell 141, 1117–1134  c The Authors Journal compilation  c 2011 Biochemical Society

22 Kappert, K., Peters, K. G., Bohmer, F. D. and Ostman, A. (2005) Tyrosine phosphatases in vessel wall signaling. Cardiovasc. Res. 65, 587–598 23 de Vries, C., Escobedo, J. A., Ueno, H., Houck, K., Ferrara, N. and Williams, L. T. (1992) The fms-like tyrosine kinase, a receptor for vascular endothelial growth factor. Science 255, 989–991 24 Shibuya, M., Yamaguchi, S., Yamane, A., Ikeda, T., Tojo, A., Matsushime, H. and Sato, M. (1990) Nucleotide sequence and expression of a novel human receptor-type tyrosine kinase gene (flt ) closely related to the fms family. Oncogene 5, 519–524 25 Davis-Smyth, T., Chen, H., Park, J., Presta, L. G. and Ferrara, N. (1996) The second immunoglobulin-like domain of the VEGF tyrosine kinase receptor Flt-1 determines ligand binding and may initiate a signal transduction cascade. EMBO J. 15, 4919–4927 26 Tanaka, K., Yamaguchi, S., Sawano, A. and Shibuya, M. (1997) Characterization of the extracellular domain in vascular endothelial growth factor receptor-1 (Flt-1 tyrosine kinase). Jpn. J. Cancer Res. 88, 867–876 27 Wiesmann, C., Fuh, G., Christinger, H. W., Eigenbrot, C., Wells, J. A. and de Vos, A. M. (1997) Crystal structure at 1.7 A˚ resolution of VEGF in complex with domain 2 of the Flt-1 receptor. Cell 91, 695–704 28 Christinger, H. W., Fuh, G., de Vos, A. M. and Wiesmann, C. (2004) The crystal structure of placental growth factor in complex with domain 2 of vascular endothelial growth factor receptor-1. J. Biol. Chem. 279, 10382–10388 29 Peters, K. G., De Vries, C. and Williams, L. T. (1993) Vascular endothelial growth factor receptor expression during embryogenesis and tissue repair suggests a role in endothelial differentiation and blood vessel growth. Proc. Natl. Acad. Sci. U.S.A. 90, 8915–8919 30 Jakeman, L. B., Winer, J., Bennett, G. L., Altar, C. A. and Ferrara, N. (1992) Binding sites for vascular endothelial growth factor are localized on endothelial cells in adult rat tissues. J. Clin. Invest. 89, 244–253 31 Clark, D. E., Smith, S. K., Sharkey, A. M. and Charnock-Jones, D. S. (1996) Localization of VEGF and expression of its receptors flt and KDR in human placenta throughout pregnancy. Hum. Reprod. 11, 1090–1098 32 Grosskreutz, C. L., Anand-Apte, B., Duplaa, C., Quinn, T. P., Terman, B. I., Zetter, B. and D’Amore, P. A. (1999) Vascular endothelial growth factor-induced migration of vascular smooth muscle cells in vitro . Microvasc. Res. 58, 128–136 33 Sawano, A., Iwai, S., Sakurai, Y., Ito, M., Shitara, K., Nakahata, T. and Shibuya, M. (2001) Flt-1, vascular endothelial growth factor receptor 1, is a novel cell surface marker for the lineage of monocyte-macrophages in humans. Blood 97, 785–791 34 Schwartz, J. D., Rowinsky, E. K., Youssoufian, H., Pytowski, B. and Wu, Y. (2010) Vascular endothelial growth factor receptor-1 in human cancer: concise review and rationale for development of IMC-18F1 (human antibody targeting vascular endothelial growth factor receptor-1). Cancer 116, 1027–1032 35 Takahashi, T., Shirasawa, T., Miyake, K., Yahagi, Y., Maruyama, N., Kasahara, N., Kawamura, T., Matsumura, O., Mitarai, T. and Sakai, O. (1995) Protein tyrosine kinases expressed in glomeruli and cultured glomerular cells: Flt-1 and VEGF expression in renal mesangial cells. Biochem. Biophys. Res. Commun. 209, 218–226 36 Nomura, M., Yamagishi, S., Harada, S., Hayashi, Y., Yamashima, T., Yamashita, J. and Yamamoto, H. (1995) Possible participation of autocrine and paracrine vascular endothelial growth factors in hypoxia-induced proliferation of endothelial cells and pericytes. J. Biol. Chem. 270, 28316–28324 37 Kendall, R. L. and Thomas, K. A. (1993) Inhibition of vascular endothelial cell growth factor activity by an endogenously encoded soluble receptor. Proc. Natl. Acad. Sci. U.S.A. 90, 10705–10709 38 Thomas, C. P., Raikwar, N. S., Kelley, E. A. and Liu, K. Z. (2010) Alternate processing of Flt1 transcripts is directed by conserved cis -elements within an intronic region of FLT1 that reciprocally regulates splicing and polyadenylation. Nucleic Acids Res. 38, 5130–5140 39 Mezquita, B., Mezquita, J., Pau, M. and Mezquita, C. (2010) A novel intracellular isoform of VEGFR-1 activates Src and promotes cell invasion in MDA-MB-231 breast cancer cells. J. Cell. Biochem. 110, 732–742 40 Shinkai, A., Ito, M., Anazawa, H., Yamaguchi, S., Shitara, K. and Shibuya, M. (1998) Mapping of the sites involved in ligand association and dissociation at the extracellular domain of the kinase insert domain-containing receptor for vascular endothelial growth factor. J. Biol. Chem. 273, 31283–31288 41 Gille, H., Kowalski, J., Yu, L., Chen, H., Pisabarro, M. T., Davis-Smyth, T. and Ferrara, N. (2000) A repressor sequence in the juxtamembrane domain of Flt-1 (VEGFR-1) constitutively inhibits vascular endothelial growth factor-dependent phosphatidylinositol 3 -kinase activation and endothelial cell migration. EMBO J. 19, 4064–4073 42 Meyer, R. D., Mohammadi, M. and Rahimi, N. (2006) A single amino acid substitution in the activation loop defines the decoy characteristic of VEGFR-1/FLT-1. J. Biol. Chem. 281, 867–875 43 Ito, N., Wernstedt, C., Engstrom, U. and Claesson-Welsh, L. (1998) Identification of vascular endothelial growth factor receptor-1 tyrosine phosphorylation sites and binding of SH2 domain-containing molecules. J. Biol. Chem. 273, 23410–23418

Signal transduction by VEGF receptors 44 Cunningham, S. A., Arrate, M. P., Brock, T. A. and Waxham, M. N. (1997) Interactions of FLT-1 and KDR with phospholipase Cγ : identification of the phosphotyrosine binding sites. Biochem. Biophys. Res. Commun. 240, 635–639 45 Yu, Y., Hulmes, J. D., Herley, M. T., Whitney, R. G., Crabb, J. W. and Sato, J. D. (2001) Direct identification of a major autophosphorylation site on vascular endothelial growth factor receptor Flt-1 that mediates phosphatidylinositol 3 -kinase binding. Biochem. J. 358, 465–472 46 Autiero, M., Waltenberger, J., Communi, D., Kranz, A., Moons, L., Lambrechts, D., Kroll, J., Plaisance, S., De Mol, M., Bono, F. et al. (2003) Role of PlGF in the intra- and intermolecular cross talk between the VEGF receptors Flt1 and Flk1. Nat. Med. 9, 936–943 47 Sawano, A., Takahashi, T., Yamaguchi, S. and Shibuya, M. (1997) The phosphorylated 1169-tyrosine containing region of flt-1 kinase (VEGFR-1) is a major binding site for PLCγ . Biochem. Biophys. Res. Commun. 238, 487–491 48 Igarashi, K., Isohara, T., Kato, T., Shigeta, K., Yamano, T. and Uno, I. (1998) Tyrosine 1213 of Flt-1 is a major binding site of Nck and SHP-2. Biochem. Biophys. Res. Commun. 246, 95–99 49 Cunningham, S. A., Waxham, M. N., Arrate, P. M. and Brock, T. A. (1995) Interaction of the Flt-1 tyrosine kinase receptor with the p85 subunit of phosphatidylinositol 3-kinase. Mapping of a novel site involved in binding. J. Biol. Chem. 270, 20254–20257 50 Gille, H., Kowalski, J., Li, B., LeCouter, J., Moffat, B., Zioncheck, T. F., Pelletier, N. and Ferrara, N. (2001) Analysis of biological effects and signaling properties of Flt-1 (VEGFR-1) and KDR (VEGFR-2). A reassessment using novel receptor-specific vascular endothelial growth factor mutants. J. Biol. Chem. 276, 3222–3230 51 Rahimi, N., Dayanir, V. and Lashkari, K. (2000) Receptor chimeras indicate that the vascular endothelial growth factor receptor-1 (VEGFR-1) modulates mitogenic activity of VEGFR-2 in endothelial cells. J. Biol. Chem. 275, 16986–16992 52 Waltenberger, J., Claesson-Welsh, L., Siegbahn, A., Shibuya, M. and Heldin, C. H. (1994) Different signal transduction properties of KDR and Flt1, two receptors for vascular endothelial growth factor. J. Biol. Chem. 269, 26988–26995 53 Kanno, S., Oda, N., Abe, M., Terai, Y., Ito, M., Shitara, K., Tabayashi, K., Shibuya, M. and Sato, Y. (2000) Roles of two VEGF receptors, Flt-1 and KDR, in the signal transduction of VEGF effects in human vascular endothelial cells. Oncogene 19, 2138–2146 54 Wang, F., Yamauchi, M., Muramatsu, M., Osawa, T., Tsuchida, R. and Shibuya, M. (2011) RACK1 regulates VEGF/Flt1-mediated cell migration via activation of a PI3-K/Akt pathway. J. Biol. Chem. 286, 9097–9106 55 Cai, J., Ahmad, S., Jiang, W. G., Huang, J., Kontos, C. D., Boulton, M. and Ahmed, A. (2003) Activation of vascular endothelial growth factor receptor-1 sustains angiogenesis and Bcl-2 expression via the phosphatidylinositol 3-kinase pathway in endothelial cells. Diabetes 52, 2959–2968 56 Huang, K., Andersson, C., Roomans, G. M., Ito, N. and Claesson-Welsh, L. (2001) Signaling properties of VEGF receptor-1 and -2 homo- and heterodimers. Int. J. Biochem. Cell Biol. 33, 315–324 57 Mac Gabhann, F. and Popel, A. S. (2007) Dimerization of VEGF receptors and implications for signal transduction: a computational study. Biophys. Chem. 128, 125–139 58 Kou, R., SenBanerjee, S., Jain, M. K. and Michel, T. (2005) Differential regulation of vascular endothelial growth factor receptors (VEGFR) revealed by RNA interference: interactions of VEGFR-1 and VEGFR-2 in endothelial cell signaling. Biochemistry 44, 15064–15073 59 Jinnin, M., Medici, D., Park, L., Limaye, N., Liu, Y., Boscolo, E., Bischoff, J., Vikkula, M., Boye, E. and Olsen, B. R. (2008) Suppressed NFAT-dependent VEGFR1 expression and constitutive VEGFR2 signaling in infantile hemangioma. Nat. Med. 14, 1236–1246 60 Ding, Y., Huang, Y., Song, N., Gao, X., Yuan, S., Wang, X., Cai, H., Fu, Y. and Luo, Y. (2010) NFAT1 mediates placental growth factor-induced myelomonocytic cell recruitment via the induction of TNF-α. J. Immunol. 184, 2593–2601 61 Tchaikovski, V., Fellbrich, G. and Waltenberger, J. (2008) The molecular basis of VEGFR-1 signal transduction pathways in primary human monocytes. Arterioscler. Thromb. Vasc. Biol. 28, 322–328 62 Bellik, L., Vinci, M. C., Filippi, S., Ledda, F. and Parenti, A. (2005) Intracellular pathways triggered by the selective FLT-1-agonist placental growth factor in vascular smooth muscle cells exposed to hypoxia. Br. J. Pharmacol. 146, 568–575 63 Lesslie, D. P., Summy, J. M., Parikh, N. U., Fan, F., Trevino, J. G., Sawyer, T. K., Metcalf, C. A., Shakespeare, W. C., Hicklin, D. J., Ellis, L. M. and Gallick, G. E. (2006) Vascular endothelial growth factor receptor-1 mediates migration of human colorectal carcinoma cells by activation of Src family kinases. Br. J. Cancer. 94, 1710–1717 64 Taylor, A. P., Leon, E. and Goldenberg, D. M. (2010) Placental growth factor (PlGF) enhances breast cancer cell motility by mobilising ERK1/2 phosphorylation and cytoskeletal rearrangement. Br. J. Cancer. 103, 82–89 65 Fong, G. H., Rossant, J., Gertsenstein, M. and Breitman, M. L. (1995) Role of the Flt-1 receptor tyrosine kinase in regulating the assembly of vascular endothelium. Nature 376, 66–70

179

66 Hiratsuka, S., Minowa, O., Kuno, J., Noda, T. and Shibuya, M. (1998) Flt-1 lacking the tyrosine kinase domain is sufficient for normal development and angiogenesis in mice. Proc. Natl. Acad. Sci. U.S.A. 95, 9349–9354 67 Kappas, N. C., Zeng, G., Chappell, J. C., Kearney, J. B., Hazarika, S., Kallianos, K. G., Patterson, C., Annex, B. H. and Bautch, V. L. (2008) The VEGF receptor Flt-1 spatially modulates Flk-1 signaling and blood vessel branching. J. Cell Biol. 181, 847–858 68 Hiratsuka, S., Nakao, K., Nakamura, K., Katsuki, M., Maru, Y. and Shibuya, M. (2005) Membrane fixation of vascular endothelial growth factor receptor 1 ligand-binding domain is important for vasculogenesis and angiogenesis in mice. Mol. Cell. Biol. 25, 346–354 69 Kendall, R. L., Wang, G., DiSalvo, J. and Thomas, K. A. (1994) Specificity of vascular endothelial cell growth factor receptor ligand binding domains. Biochem. Biophys. Res. Commun. 201, 326–330 70 Lorquet, S., Berndt, S., Blacher, S., Gengoux, E., Peulen, O., Maquoi, E., Noel, A., Foidart, J. M., Munaut, C. and Pequeux, C. (2010) Soluble forms of VEGF receptor-1 and -2 promote vascular maturation via mural cell recruitment. FASEB J. 24, 3782–3795 71 Ambati, B. K., Nozaki, M., Singh, N., Takeda, A., Jani, P. D., Suthar, T., Albuquerque, R. J., Richter, E., Sakurai, E., Newcomb, M. T. et al. (2006) Corneal avascularity is due to soluble VEGF receptor-1. Nature 443, 993–997 72 Clark, D. E., Smith, S. K., He, Y., Day, K. A., Licence, D. R., Corps, A. N., Lammoglia, R. and Charnock-Jones, D. S. (1998) A vascular endothelial growth factor antagonist is produced by the human placenta and released into the maternal circulation. Biol. Reprod. 59, 1540–1548 73 Maynard, S. E., Min, J. Y., Merchan, J., Lim, K. H., Li, J., Mondal, S., Libermann, T. A., Morgan, J. P., Sellke, F. W., Stillman, I. E. et al. (2003) Excess placental soluble fms-like tyrosine kinase 1 (sFlt1) may contribute to endothelial dysfunction, hypertension, and proteinuria in preeclampsia. J. Clin. Invest. 111, 649–658 74 Toi, M., Bando, H., Ogawa, T., Muta, M., Hornig, C. and Weich, H. A. (2002) Significance of vascular endothelial growth factor (VEGF)/soluble VEGF receptor-1 relationship in breast cancer. Int. J. Cancer. 98, 14–18 75 Scheufler, K. M., Drevs, J., van Velthoven, V., Reusch, P., Klisch, J., Augustin, H. G., Zentner, J. and Marme, D. (2003) Implications of vascular endothelial growth factor, sFlt-1, and sTie-2 in plasma, serum and cerebrospinal fluid during cerebral ischemia in man. J. Cereb. Blood Flow Metab. 23, 99–110 76 Hiratsuka, S., Maru, Y., Okada, A., Seiki, M., Noda, T. and Shibuya, M. (2001) Involvement of Flt-1 tyrosine kinase (vascular endothelial growth factor receptor-1) in pathological angiogenesis. Cancer Res. 61, 1207–1213 77 Hiratsuka, S., Nakamura, K., Iwai, S., Murakami, M., Itoh, T., Kijima, H., Shipley, J. M., Senior, R. M. and Shibuya, M. (2002) MMP9 induction by vascular endothelial growth factor receptor-1 is involved in lung-specific metastasis. Cancer Cell 2, 289–300 78 Kami, J., Muranaka, K., Yanagi, Y., Obata, R., Tamaki, Y. and Shibuya, M. (2008) Inhibition of choroidal neovascularization by blocking vascular endothelial growth factor receptor tyrosine kinase. Jpn. J. Ophthalmol. 52, 91–98 79 Murakami, M., Iwai, S., Hiratsuka, S., Yamauchi, M., Nakamura, K., Iwakura, Y. and Shibuya, M. (2006) Signaling of vascular endothelial growth factor receptor-1 tyrosine kinase promotes rheumatoid arthritis through activation of monocytes/macrophages. Blood 108, 1849–1856 80 Beck, H., Raab, S., Copanaki, E., Heil, M., Scholz, A., Shibuya, M., Deller, T., Machein, M. and Plate, K. H. (2010) VEGFR-1 signaling regulates the homing of bone marrow-derived cells in a mouse stroke model. J. Neuropathol. Exp. Neurol. 69, 168–175 81 Carmeliet, P., Moons, L., Luttun, A., Vincenti, V., Compernolle, V., De Mol, M., Wu, Y., Bono, F., Devy, L., Beck, H. et al. (2001) Synergism between vascular endothelial growth factor and placental growth factor contributes to angiogenesis and plasma extravasation in pathological conditions. Nat. Med. 7, 575–583 82 Fischer, C., Mazzone, M., Jonckx, B. and Carmeliet, P. (2008) FLT1 and its ligands VEGFB and PlGF: drug targets for anti-angiogenic therapy? Nat. Rev. Cancer. 8, 942–956 83 Li, X., Tjwa, M., Van Hove, I., Enholm, B., Neven, E., Paavonen, K., Jeltsch, M., Juan, T. D., Sievers, R. E., Chorianopoulos, E. et al. (2008) Reevaluation of the role of VEGF-B suggests a restricted role in the revascularization of the ischemic myocardium. Arterioscler. Thromb. Vasc. Biol. 28, 1614–1620 84 Hagberg, C. E., Falkevall, A., Wang, X., Larsson, E., Huusko, J., Nilsson, I., van Meeteren, L. A., Samen, E., Lu, L., Vanwildemeersch, M. et al. (2010) Vascular endothelial growth factor B controls endothelial fatty acid uptake. Nature 464, 917–921 85 Aase, K., von Euler, G., Li, X., Ponten, A., Thoren, P., Cao, R., Cao, Y., Olofsson, B., Gebre-Medhin, S., Pekny, M. et al. (2001) Vascular endothelial growth factor-B-deficient mice display an atrial conduction defect. Circulation 104, 358–364 86 Bellomo, D., Headrick, J. P., Silins, G. U., Paterson, C. A., Thomas, P. S., Gartside, M., Mould, A., Cahill, M. M., Tonks, I. D., Grimmond, S. M. et al. (2000) Mice lacking the vascular endothelial growth factor-B gene (Vegfb ) have smaller hearts, dysfunctional coronary vasculature, and impaired recovery from cardiac ischemia. Circ. Res. 86, E29–E35  c The Authors Journal compilation  c 2011 Biochemical Society

180

S. Koch and others

87 Kobayashi, S., Sawano, A., Nojima, Y., Shibuya, M. and Maru, Y. (2004) The c-Cbl/CD2AP complex regulates VEGF-induced endocytosis and degradation of Flt-1 (VEGFR-1). FASEB J. 18, 929–931 88 Fuh, G., Li, B., Crowley, C., Cunningham, B. and Wells, J. A. (1998) Requirements for binding and signaling of the kinase domain receptor for vascular endothelial growth factor. J. Biol. Chem. 273, 11197–11204 89 McColl, B. K., Baldwin, M. E., Roufail, S., Freeman, C., Moritz, R. L., Simpson, R. J., Alitalo, K., Stacker, S. A. and Achen, M. G. (2003) Plasmin activates the lymphangiogenic growth factors VEGF-C and VEGF-D. J. Exp. Med. 198, 863–868 90 Albuquerque, R. J., Hayashi, T., Cho, W. G., Kleinman, M. E., Dridi, S., Takeda, A., Baffi, J. Z., Yamada, K., Kaneko, H., Green, M. G. et al. (2009) Alternatively spliced vascular endothelial growth factor receptor-2 is an essential endogenous inhibitor of lymphatic vessel growth. Nat. Med. 15, 1023–1030 91 Pavlakovic, H., Becker, J., Albuquerque, R., Wilting, J. and Ambati, J. (2010) Soluble VEGFR-2: an antilymphangiogenic variant of VEGF receptors. Ann. N.Y. Acad. Sci. 1207 (Suppl. 1), E7–E15 92 Becker, J., Pavlakovic, H., Ludewig, F., Wilting, F., Weich, H. A., Albuquerque, R., Ambati, J. and Wilting, J. (2010) Neuroblastoma progression correlates with downregulation of the lymphangiogenesis inhibitor sVEGFR-2. Clin. Cancer Res. 16, 1431–1441 93 Millauer, B., Wizigmann-Voos, S., Schnurch, H., Martinez, R., Moller, N. P., Risau, W. and Ullrich, A. (1993) High affinity VEGF binding and developmental expression suggest Flk-1 as a major regulator of vasculogenesis and angiogenesis. Cell 72, 835–846 94 Quinn, T. P., Peters, K. G., De Vries, C., Ferrara, N. and Williams, L. T. (1993) Fetal liver kinase 1 is a receptor for vascular endothelial growth factor and is selectively expressed in vascular endothelium. Proc. Natl. Acad. Sci. U.S.A. 90, 7533–7537 95 Oelrichs, R. B., Reid, H. H., Bernard, O., Ziemiecki, A. and Wilks, A. F. (1993) NYK/FLK-1: a putative receptor protein tyrosine kinase isolated from E10 embryonic neuroepithelium is expressed in endothelial cells of the developing embryo. Oncogene 8, 11–18 96 Matsumoto, T. and Claesson-Welsh, L. (2001) VEGF receptor signal transduction. Sci. STKE 2001, re21 97 Tasaki, Y., Nishimura, R., Shibaya, M., Lee, H. Y., Acosta, T. J. and Okuda, K. (2010) Expression of VEGF and its receptors in the bovine endometrium throughout the estrous cycle: effects of VEGF on prostaglandin production in endometrial cells. J. Reprod. Dev. 56, 223–229 98 Millauer, B., Shawver, L. K., Plate, K. H., Risau, W. and Ullrich, A. (1994) Glioblastoma growth inhibited in vivo by a dominant-negative Flk-1 mutant. Nature 367, 576–579 99 Plate, K. H., Breier, G., Millauer, B., Ullrich, A. and Risau, W. (1993) Up-regulation of vascular endothelial growth factor and its cognate receptors in a rat glioma model of tumor angiogenesis. Cancer Res. 53, 5822–5827 100 Youssoufian, H., Hicklin, D. J. and Rowinsky, E. K. (2007) Review: monoclonal antibodies to the vascular endothelial growth factor receptor-2 in cancer therapy. Clin. Cancer Res. 13, 5544s–5548s 101 Maharaj, A. S., Saint-Geniez, M., Maldonado, A. E. and D’Amore, P. A. (2006) Vascular endothelial growth factor localization in the adult. Am. J. Pathol. 168, 639–648 102 Matsumoto, T., Bohman, S., Dixelius, J., Berge, T., Dimberg, A., Magnusson, P., Wang, L., Wikner, C., Qi, J. H., Wernstedt, C. et al. (2005) VEGF receptor-2 Y951 signaling and a role for the adapter molecule TSAd in tumor angiogenesis. EMBO J. 24, 2342–2353 103 Takahashi, T., Yamaguchi, S., Chida, K. and Shibuya, M. (2001) A single autophosphorylation site on KDR/Flk-1 is essential for VEGF-A-dependent activation of PLC-γ and DNA synthesis in vascular endothelial cells. EMBO J. 20, 2768–2778 104 Solowiej, J., Bergqvist, S., McTigue, M. A., Marrone, T., Quenzer, T., Cobbs, M., Ryan, K., Kania, R. S., Diehl, W. and Murray, B. W. (2009) Characterizing the effects of the juxtamembrane domain on vascular endothelial growth factor receptor-2 enzymatic activity, autophosphorylation, and inhibition by axitinib. Biochemistry 48, 7019–7031 105 Wu, L. W., Mayo, L. D., Dunbar, J. D., Kessler, K. M., Ozes, O. N., Warren, R. S. and Donner, D. B. (2000) VRAP is an adaptor protein that binds KDR, a receptor for vascular endothelial cell growth factor. J. Biol. Chem. 275, 6059–6062 106 Kendall, R. L., Rutledge, R. Z., Mao, X., Tebben, A. J., Hungate, R. W. and Thomas, K. A. (1999) Vascular endothelial growth factor receptor KDR tyrosine kinase activity is increased by autophosphorylation of two activation loop tyrosine residues. J. Biol. Chem. 274, 6453–6460 107 Meyer, R. D., Sacks, D. B. and Rahimi, N. (2008) IQGAP1-dependent signaling pathway regulates endothelial cell proliferation and angiogenesis. PLoS ONE 3, e3848 108 Yamaoka-Tojo, M., Tojo, T., Kim, H. W., Hilenski, L., Patrushev, N. A., Zhang, L., Fukai, T. and Ushio-Fukai, M. (2006) IQGAP1 mediates VE-cadherin-based cell–cell contacts and VEGF signaling at adherence junctions linked to angiogenesis. Arterioscler. Thromb. Vasc. Biol. 26, 1991–1997 109 Holmqvist, K., Cross, M. J., Rolny, C., Hagerkvist, R., Rahimi, N., Matsumoto, T., Claesson-Welsh, L. and Welsh, M. (2004) The adaptor protein shb binds to tyrosine 1175 in vascular endothelial growth factor (VEGF) receptor-2 and regulates VEGF-dependent cellular migration. J. Biol. Chem. 279, 22267–22275  c The Authors Journal compilation  c 2011 Biochemical Society

110 Warner, A. J., Lopez-Dee, J., Knight, E. L., Feramisco, J. R. and Prigent, S. A. (2000) The Shc-related adaptor protein, Sck, forms a complex with the vascular-endothelial-growth-factor receptor KDR in transfected cells. Biochem. J. 347, 501–509 111 Abu-Ghazaleh, R., Kabir, J., Jia, H., Lobo, M. and Zachary, I. (2001) Src mediates stimulation by vascular endothelial growth factor of the phosphorylation of focal adhesion kinase at tyrosine 861, and migration and anti-apoptosis in endothelial cells. Biochem. J. 360, 255–264 112 Holmqvist, K., Cross, M., Riley, D. and Welsh, M. (2003) The Shb adaptor protein causes Src-dependent cell spreading and activation of focal adhesion kinase in murine brain endothelial cells. Cell. Signalling 15, 171–179 113 Kroll, J. and Waltenberger, J. (1997) The vascular endothelial growth factor receptor KDR activates multiple signal transduction pathways in porcine aortic endothelial cells. J. Biol. Chem. 272, 32521–32527 114 Meadows, K. N., Bryant, P. and Pumiglia, K. (2001) Vascular endothelial growth factor induction of the angiogenic phenotype requires Ras activation. J. Biol. Chem. 276, 49289–49298 115 Zachary, I. (2001) Signaling mechanisms mediating vascular protective actions of vascular endothelial growth factor. Am. J. Physiol. 280, C1375–C1386 116 Takahashi, T., Ueno, H. and Shibuya, M. (1999) VEGF activates protein kinase C-dependent, but Ras-independent Raf-MEK-MAP kinase pathway for DNA synthesis in primary endothelial cells. Oncogene 18, 2221–2230 117 Xia, P., Aiello, L. P., Ishii, H., Jiang, Z. Y., Park, D. J., Robinson, G. S., Takagi, H., Newsome, W. P., Jirousek, M. R. and King, G. L. (1996) Characterization of vascular endothelial growth factor’s effect on the activation of protein kinase C, its isoforms, and endothelial cell growth. J. Clin. Invest. 98, 2018–2026 118 Wellner, M., Maasch, C., Kupprion, C., Lindschau, C., Luft, F. C. and Haller, H. (1999) The proliferative effect of vascular endothelial growth factor requires protein kinase C-α and protein kinase C-ξ . Arterioscler. Thromb. Vasc. Biol. 19, 178–185 119 Shu, X., Wu, W., Mosteller, R. D. and Broek, D. (2002) Sphingosine kinase mediates vascular endothelial growth factor-induced activation of ras and mitogen-activated protein kinases. Mol. Cell. Biol. 22, 7758–7768 120 Bhattacharya, R., Kwon, J., Li, X., Wang, E., Patra, S., Bida, J. P., Bajzer, Z., Claesson-Welsh, L. and Mukhopadhyay, D. (2009) Distinct role of PLCβ3 in VEGF-mediated directional migration and vascular sprouting. J. Cell Sci. 122, 1025–1034 121 Wong, C. and Jin, Z. G. (2005) Protein kinase C-dependent protein kinase D activation modulates ERK signal pathway and endothelial cell proliferation by vascular endothelial growth factor. J. Biol. Chem. 280, 33262–33269 122 Wang, S., Li, X., Parra, M., Verdin, E., Bassel-Duby, R. and Olson, E. N. (2008) Control of endothelial cell proliferation and migration by VEGF signaling to histone deacetylase 7. Proc. Natl. Acad. Sci. U.S.A. 105, 7738–7743 123 Ha, C. H., Wang, W., Jhun, B. S., Wong, C., Hausser, A., Pfizenmaier, K., McKinsey, T. A., Olson, E. N. and Jin, Z. G. (2008) Protein kinase D-dependent phosphorylation and nuclear export of histone deacetylase 5 mediates vascular endothelial growth factor-induced gene expression and angiogenesis. J. Biol. Chem. 283, 14590–14599 124 Evans, I. M., Britton, G. and Zachary, I. C. (2008) Vascular endothelial growth factor induces heat shock protein (HSP) 27 serine 82 phosphorylation and endothelial tubulogenesis via protein kinase D and independent of p38 kinase. Cell. Signalling 20, 1375–1384 125 Lamalice, L., Houle, F. and Huot, J. (2006) Phosphorylation of Tyr1214 within VEGFR-2 triggers the recruitment of Nck and activation of Fyn leading to SAPK2/p38 activation and endothelial cell migration in response to VEGF. J. Biol. Chem. 281, 34009–34020 126 Lamalice, L., Houle, F., Jourdan, G. and Huot, J. (2004) Phosphorylation of tyrosine 1214 on VEGFR2 is required for VEGF-induced activation of Cdc42 upstream of SAPK2/p38. Oncogene 23, 434–445 127 Laramee, M., Chabot, C., Cloutier, M., Stenne, R., Holgado-Madruga, M., Wong, A. J. and Royal, I. (2007) The scaffolding adapter Gab1 mediates vascular endothelial growth factor signaling and is required for endothelial cell migration and capillary formation. J. Biol. Chem. 282, 7758–7769 128 Jin, Z. G., Wong, C., Wu, J. and Berk, B. C. (2005) Flow shear stress stimulates Gab1 tyrosine phosphorylation to mediate protein kinase B and endothelial nitric-oxide synthase activation in endothelial cells. J. Biol. Chem. 280, 12305–12309 129 Datta, S. R., Brunet, A. and Greenberg, M. E. (1999) Cellular survival: a play in three Akts. Genes Dev. 13, 2905–2927 130 Los, M., Panigrahi, S., Rashedi, I., Mandal, S., Stetefeld, J., Essmann, F. and Schulze-Osthoff, K. (2009) Apoptin, a tumor-selective killer. Biochim. Biophys. Acta 1793, 1335–1342 131 Gerber, H. P., McMurtrey, A., Kowalski, J., Yan, M., Keyt, B. A., Dixit, V. and Ferrara, N. (1998) Vascular endothelial growth factor regulates endothelial cell survival through the phosphatidylinositol 3 -kinase/Akt signal transduction pathway. Requirement for Flk-1/KDR activation. J. Biol. Chem. 273, 30336–30343

Signal transduction by VEGF receptors 132 Kumar, P., Miller, A. I. and Polverini, P. J. (2004) p38 MAPK mediates γ -irradiation-induced endothelial cell apoptosis, and vascular endothelial growth factor protects endothelial cells through the phosphoinositide 3-kinase–Akt–Bcl-2 pathway. J. Biol. Chem. 279, 43352–43360 133 Tran, J., Rak, J., Sheehan, C., Saibil, S. D., LaCasse, E., Korneluk, R. G. and Kerbel, R. S. (1999) Marked induction of the IAP family antiapoptotic proteins survivin and XIAP by VEGF in vascular endothelial cells. Biochem. Biophys. Res. Commun. 264, 781–788 134 Deveraux, Q. L., Roy, N., Stennicke, H. R., Van Arsdale, T., Zhou, Q., Srinivasula, S. M., Alnemri, E. S., Salvesen, G. S. and Reed, J. C. (1998) IAPs block apoptotic events induced by caspase-8 and cytochrome c by direct inhibition of distinct caspases. EMBO J. 17, 2215–2223 135 Shalaby, F., Rossant, J., Yamaguchi, T. P., Gertsenstein, M., Wu, X. F., Breitman, M. L. and Schuh, A. C. (1995) Failure of blood-island formation and vasculogenesis in Flk-1-deficient mice. Nature 376, 62–66 136 Carmeliet, P., Ferreira, V., Breier, G., Pollefeyt, S., Kieckens, L., Gertsenstein, M., Fahrig, M., Vandenhoeck, A., Harpal, K., Eberhardt, C. et al. (1996) Abnormal blood vessel development and lethality in embryos lacking a single VEGF allele. Nature 380, 435–439 137 Ferrara, N., Carver-Moore, K., Chen, H., Dowd, M., Lu, L., O’Shea, K. S., Powell-Braxton, L., Hillan, K. J. and Moore, M. W. (1996) Heterozygous embryonic lethality induced by targeted inactivation of the VEGF gene. Nature 380, 439–442 138 Sakurai, Y., Ohgimoto, K., Kataoka, Y., Yoshida, N. and Shibuya, M. (2005) Essential role of Flk-1 (VEGF receptor 2) tyrosine residue 1173 in vasculogenesis in mice. Proc. Natl. Acad. Sci. U.S.A. 102, 1076–1081 139 Ji, Q. S., Winnier, G. E., Niswender, K. D., Horstman, D., Wisdom, R., Magnuson, M. A. and Carpenter, G. (1997) Essential role of the tyrosine kinase substrate phospholipase C-gamma1 in mammalian growth and development. Proc. Natl. Acad. Sci. U.S.A. 94, 2999–3003 140 Liao, H. J., Kume, T., McKay, C., Xu, M. J., Ihle, J. N. and Carpenter, G. (2002) Absence of erythrogenesis and vasculogenesis in Plcg1-deficient mice. J. Biol. Chem. 277, 9335–9341 141 Lawson, N. D., Mugford, J. W., Diamond, B. A. and Weinstein, B. M. (2003) Phospholipase Cγ -1 is required downstream of vascular endothelial growth factor during arterial development. Genes Dev. 17, 1346–1351 142 Funa, N. S., Kriz, V., Zang, G., Calounova, G., Akerblom, B., Mares, J., Larsson, E., Sun, Y., Betsholtz, C. and Welsh, M. (2009) Dysfunctional microvasculature as a consequence of shb gene inactivation causes impaired tumor growth. Cancer Res. 69, 2141–2148 143 Esser, S., Wolburg, K., Wolburg, H., Breier, G., Kurzchalia, T. and Risau, W. (1998) Vascular endothelial growth factor induces endothelial fenestrations in vitro . J. Cell Biol. 140, 947–959 144 Bates, D. O. and Harper, S. J. (2003) Regulation of vascular permeability by vascular endothelial growth factors. Vasc. Pharmacol. 39, 225–237 145 Nagy, J. A., Benjamin, L., Zeng, H., Dvorak, A. M. and Dvorak, H. F. (2008) Vascular permeability, vascular hyperpermeability and angiogenesis. Angiogenesis 11, 109–119 146 Garrido-Urbani, S., Bradfield, P. F., Lee, B. P. and Imhof, B. A. (2008) Vascular and epithelial junctions: a barrier for leucocyte migration. Biochem. Soc. Trans. 36, 203–211 147 Nagy, J. A., Vasile, E., Feng, D., Sundberg, C., Brown, L. F., Manseau, E. J., Dvorak, A. M. and Dvorak, H. F. (2002) VEGF-A induces angiogenesis, arteriogenesis, lymphangiogenesis, and vascular malformations. Cold Spring Harb. Symp. Quant. Biol. 67, 227–237 148 Labrecque, L., Royal, I., Surprenant, D. S., Patterson, C., Gingras, D. and Beliveau, R. (2003) Regulation of vascular endothelial growth factor receptor-2 activity by caveolin-1 and plasma membrane cholesterol. Mol. Biol. Cell 14, 334–347 149 Lin, M. I., Yu, J., Murata, T. and Sessa, W. C. (2007) Caveolin-1-deficient mice have increased tumor microvascular permeability, angiogenesis, and growth. Cancer Res. 67, 2849–2856 150 Dejana, E. (2004) Endothelial cell–cell junctions: happy together. Nat. Rev. Mol. Cell Biol. 5, 261–270 151 Rudini, N. and Dejana, E. (2008) Adherens junctions. Curr. Biol. 18, R1080–R1082 152 Eliceiri, B. P., Paul, R., Schwartzberg, P. L., Hood, J. D., Leng, J. and Cheresh, D. A. (1999) Selective requirement for Src kinases during VEGF-induced angiogenesis and vascular permeability. Mol. Cell 4, 915–924 153 Wallez, Y., Cand, F., Cruzalegui, F., Wernstedt, C., Souchelnytskyi, S., Vilgrain, I. and Huber, P. (2007) Src kinase phosphorylates vascular endothelial-cadherin in response to vascular endothelial growth factor: identification of tyrosine 685 as the unique target site. Oncogene 26, 1067–1077 154 Weis, S., Shintani, S., Weber, A., Kirchmair, R., Wood, M., Cravens, A., McSharry, H., Iwakura, A., Yoon, Y. S., Himes, N. et al. (2004) Src blockade stabilizes a Flk/cadherin complex, reducing edema and tissue injury following myocardial infarction. J. Clin. Invest. 113, 885–894 155 Huber, A. H. and Weis, W. I. (2001) The structure of the β-catenin/E-cadherin complex and the molecular basis of diverse ligand recognition by β-catenin. Cell 105, 391–402

181

156 Roura, S., Miravet, S., Piedra, J., Garcia de Herreros, A. and Dunach, M. (1999) Regulation of E-cadherin/catenin association by tyrosine phosphorylation. J. Biol. Chem. 274, 36734–36740 157 Esser, S., Lampugnani, M. G., Corada, M., Dejana, E. and Risau, W. (1998) Vascular endothelial growth factor induces VE-cadherin tyrosine phosphorylation in endothelial cells. J. Cell Sci. 111, 1853–1865 158 Nawroth, R., Poell, G., Ranft, A., Kloep, S., Samulowitz, U., Fachinger, G., Golding, M., Shima, D. T., Deutsch, U. and Vestweber, D. (2002) VE-PTP and VE-cadherin ectodomains interact to facilitate regulation of phosphorylation and cell contacts. EMBO J. 21, 4885–4895 159 Nottebaum, A. F., Cagna, G., Winderlich, M., Gamp, A. C., Linnepe, R., Polaschegg, C., Filippova, K., Lyck, R., Engelhardt, B., Kamenyeva, O. et al. (2008) VE-PTP maintains the endothelial barrier via plakoglobin and becomes dissociated from VE-cadherin by leukocytes and by VEGF. J. Exp. Med. 205, 2929–2945 160 Gavard, J. and Gutkind, J. S. (2006) VEGF controls endothelial-cell permeability by promoting the β-arrestin-dependent endocytosis of VE-cadherin. Nat. Cell Biol. 8, 1223–1234 161 Fukumura, D., Gohongi, T., Kadambi, A., Izumi, Y., Ang, J., Yun, C. O., Buerk, D. G., Huang, P. L. and Jain, R. K. (2001) Predominant role of endothelial nitric oxide synthase in vascular endothelial growth factor-induced angiogenesis and vascular permeability. Proc. Natl. Acad. Sci. U.S.A. 98, 2604–2609 162 Dimmeler, S., Fleming, I., Fisslthaler, B., Hermann, C., Busse, R. and Zeiher, A. M. (1999) Activation of nitric oxide synthase in endothelial cells by Akt-dependent phosphorylation. Nature 399, 601–605 163 Fulton, D., Gratton, J. P., McCabe, T. J., Fontana, J., Fujio, Y., Walsh, K., Franke, T. F., Papapetropoulos, A. and Sessa, W. C. (1999) Regulation of endothelium-derived nitric oxide production by the protein kinase Akt. Nature 399, 597–601 164 Murohara, T., Horowitz, J. R., Silver, M., Tsurumi, Y., Chen, D., Sullivan, A. and Isner, J. M. (1998) Vascular endothelial growth factor/vascular permeability factor enhances vascular permeability via nitric oxide and prostacyclin. Circulation 97, 99–107 165 Wu, H. M., Huang, Q., Yuan, Y. and Granger, H. J. (1996) VEGF induces NO-dependent hyperpermeability in coronary venules. Am. J. Physiol. 271, H2735–H2739 166 Thibeault, S., Rautureau, Y., Oubaha, M., Faubert, D., Wilkes, B. C., Delisle, C. and Gratton, J. P. (2010) S -nitrosylation of β-catenin by eNOS-derived NO promotes VEGF-induced endothelial cell permeability. Mol. Cell 39, 468–476 167 Feng, X., Gaeta, M. L., Madge, L. A., Yang, J. H., Bradley, J. R. and Pober, J. S. (2001) Caveolin-1 associates with TRAF2 to form a complex that is recruited to tumor necrosis factor receptors. J. Biol. Chem. 276, 8341–8349 168 Salikhova, A., Wang, L., Lanahan, A. A., Liu, M., Simons, M., Leenders, W. P., Mukhopadhyay, D. and Horowitz, A. (2008) Vascular endothelial growth factor and semaphorin induce neuropilin-1 endocytosis via separate pathways. Circ. Res. 103, e71–e79 169 Bauer, P. M., Yu, J., Chen, Y., Hickey, R., Bernatchez, P. N., Looft-Wilson, R., Huang, Y., Giordano, F., Stan, R. V. and Sessa, W. C. (2005) Endothelial-specific expression of caveolin-1 impairs microvascular permeability and angiogenesis. Proc. Natl. Acad. Sci. U.S.A. 102, 204–209 170 Bhattacharya, R., Kang-Decker, N., Hughes, D. A., Mukherjee, P., Shah, V., McNiven, M. A. and Mukhopadhyay, D. (2005) Regulatory role of dynamin-2 in VEGFR-2/KDR-mediated endothelial signaling. FASEB J. 19, 1692–1694 171 Lampugnani, M. G., Orsenigo, F., Gagliani, M. C., Tacchetti, C. and Dejana, E. (2006) Vascular endothelial cadherin controls VEGFR-2 internalization and signaling from intracellular compartments. J. Cell Biol. 174, 593–604 172 Sawamiphak, S., Seidel, S., Essmann, C. L., Wilkinson, G. A., Pitulescu, M. E., Acker, T. and Acker-Palmer, A. (2010) Ephrin-B2 regulates VEGFR2 function in developmental and tumour angiogenesis. Nature 465, 487–491 173 Lanahan, A. A., Hermans, K., Claes, F., Kerley-Hamilton, J. S., Zhuang, Z. W., Giordano, F. J., Carmeliet, P. and Simons, M. (2010) VEGF receptor 2 endocytic trafficking regulates arterial morphogenesis. Dev. Cell 18, 713–724 174 Ewan, L. C., Jopling, H. M., Jia, H., Mittar, S., Bagherzadeh, A., Howell, G. J., Walker, J. H., Zachary, I. C. and Ponnambalam, S. (2006) Intrinsic tyrosine kinase activity is required for vascular endothelial growth factor receptor 2 ubiquitination, sorting and degradation in endothelial cells. Traffic 7, 1270–1282 175 Gampel, A., Moss, L., Jones, M. C., Brunton, V., Norman, J. C. and Mellor, H. (2006) VEGF regulates the mobilization of VEGFR2/KDR from an intracellular endothelial storage compartment. Blood 108, 2624–2631 176 Singh, A. J., Meyer, R. D., Band, H. and Rahimi, N. (2005) The carboxyl terminus of VEGFR-2 is required for PKC-mediated down-regulation. Mol. Biol. Cell 16, 2106–2118 177 Ikeda, S., Ushio-Fukai, M., Zuo, L., Tojo, T., Dikalov, S., Patrushev, N. A. and Alexander, R. W. (2005) Novel role of ARF6 in vascular endothelial growth factor-induced signaling and angiogenesis. Circ. Res. 96, 467–475 178 Donaldson, J. G. (2002) Arf6 and its role in cytoskeletal modulation. Methods Mol. Biol. 189, 191–198  c The Authors Journal compilation  c 2011 Biochemical Society

182

S. Koch and others

179 Murdaca, J., Treins, C., Monthouel-Kartmann, M. N., Pontier-Bres, R., Kumar, S., Van Obberghen, E. and Giorgetti-Peraldi, S. (2004) Grb10 prevents Nedd4-mediated vascular endothelial growth factor receptor-2 degradation. J. Biol. Chem. 279, 26754–26761 180 Pajusola, K., Aprelikova, O., Armstrong, E., Morris, S. and Alitalo, K. (1993) Two human FLT4 receptor tyrosine kinase isoforms with distinct carboxy terminal tails are produced by alternative processing of primary transcripts. Oncogene 8, 2931–2937 181 Joukov, V., Sorsa, T., Kumar, V., Jeltsch, M., Claesson-Welsh, L., Cao, Y., Saksela, O., Kalkkinen, N. and Alitalo, K. (1997) Proteolytic processing regulates receptor specificity and activity of VEGF-C. EMBO J. 16, 3898–3911 182 Stacker, S. A., Stenvers, K., Caesar, C., Vitali, A., Domagala, T., Nice, E., Roufail, S., Simpson, R. J., Moritz, R., Karpanen, T. et al. (1999) Biosynthesis of vascular endothelial growth factor-D involves proteolytic processing which generates non-covalent homodimers. J. Biol. Chem. 274, 32127–32136 183 Makinen, T., Veikkola, T., Mustjoki, S., Karpanen, T., Catimel, B., Nice, E. C., Wise, L., Mercer, A., Kowalski, H., Kerjaschki, D. et al. (2001) Isolated lymphatic endothelial cells transduce growth, survival and migratory signals via the VEGF-C/D receptor VEGFR-3. EMBO J. 20, 4762–4773 184 Jeltsch, M., Karpanen, T., Strandin, T., Aho, K., Lankinen, H. and Alitalo, K. (2006) Vascular endothelial growth factor (VEGF)/VEGF-C mosaic molecules reveal specificity determinants and feature novel receptor binding patterns. J. Biol. Chem. 281, 12187–12195 185 Leppanen, V. M., Prota, A. E., Jeltsch, M., Anisimov, A., Kalkkinen, N., Strandin, T., Lankinen, H., Goldman, A., Ballmer-Hofer, K. and Alitalo, K. (2010) Structural determinants of growth factor binding and specificity by VEGF receptor 2. Proc. Natl. Acad. Sci. U.S.A. 107, 2425–2430 186 Baldwin, M. E., Catimel, B., Nice, E. C., Roufail, S., Hall, N. E., Stenvers, K. L., Karkkainen, M. J., Alitalo, K., Stacker, S. A. and Achen, M. G. (2001) The specificity of receptor binding by vascular endothelial growth factor-D is different in mouse and man. J. Biol. Chem. 276, 19166–19171 187 Leppanen, V. M., Jeltsch, M., Anisimov, A., Tvorogov, D., Aho, K., Kalkkinen, N., Toivanen, P., Yla-Herttuala, S., Ballmer-Hofer, K. and Alitalo, K. (2011) Structural determinants of vascular endothelial growth factor-D receptor binding and specificity. Blood 117, 1507–1515 188 Aprelikova, O., Pajusola, K., Partanen, J., Armstrong, E., Alitalo, R., Bailey, S. K., McMahon, J., Wasmuth, J., Huebner, K. and Alitalo, K. (1992) FLT4 , a novel class III receptor tyrosine kinase in chromosome 5q33-qter. Cancer Res. 52, 746–748 189 Galland, F., Karamysheva, A., Pebusque, M. J., Borg, J. P., Rottapel, R., Dubreuil, P., Rosnet, O. and Birnbaum, D. (1993) The FLT4 gene encodes a transmembrane tyrosine kinase related to the vascular endothelial growth factor receptor. Oncogene 8, 1233–1240 190 Hughes, D. C. (2001) Alternative splicing of the human VEGFGR-3 /FLT4 gene as a consequence of an integrated human endogenous retrovirus. J. Mol. Evol. 53, 77–79 191 Dixelius, J., Makinen, T., Wirzenius, M., Karkkainen, M. J., Wernstedt, C., Alitalo, K. and Claesson-Welsh, L. (2003) Ligand-induced vascular endothelial growth factor receptor-3 (VEGFR-3) heterodimerization with VEGFR-2 in primary lymphatic endothelial cells regulates tyrosine phosphorylation sites. J. Biol. Chem. 278, 40973–40979 192 Kukk, E., Lymboussaki, A., Taira, S., Kaipainen, A., Jeltsch, M., Joukov, V. and Alitalo, K. (1996) VEGF-C receptor binding and pattern of expression with VEGFR-3 suggests a role in lymphatic vascular development. Development 122, 3829–3837 193 Kaipainen, A., Korhonen, J., Mustonen, T., van Hinsbergh, V. W., Fang, G. H., Dumont, D., Breitman, M. and Alitalo, K. (1995) Expression of the fms-like tyrosine kinase 4 gene becomes restricted to lymphatic endothelium during development. Proc. Natl. Acad. Sci. U.S.A. 92, 3566–3570 194 Carmeliet, P., De Smet, F., Loges, S. and Mazzone, M. (2009) Branching morphogenesis and antiangiogenesis candidates: tip cells lead the way. Nat. Rev. Clin. Oncol. 6, 315–326 195 Tammela, T., Zarkada, G., Wallgard, E., Murtomaki, A., Suchting, S., Wirzenius, M., Waltari, M., Hellstrom, M., Schomber, T., Peltonen, R. et al. (2008) Blocking VEGFR-3 suppresses angiogenic sprouting and vascular network formation. Nature 454, 656–660 196 Siekmann, A. F. and Lawson, N. D. (2007) Notch signalling and the regulation of angiogenesis. Cell Adhes. Migr. 1, 104–106 197 Orlandini, M., Spreafico, A., Bardelli, M., Rocchigiani, M., Salameh, A., Nucciotti, S., Capperucci, C., Frediani, B. and Oliviero, S. (2006) Vascular endothelial growth factor-D activates VEGFR-3 expressed in osteoblasts inducing their differentiation. J. Biol. Chem. 281, 17961–17967 198 Le Bras, B., Barallobre, M. J., Homman-Ludiye, J., Ny, A., Wyns, S., Tammela, T., Haiko, P., Karkkainen, M. J., Yuan, L., Muriel, M. P. et al. (2006) VEGF-C is a trophic factor for neural progenitors in the vertebrate embryonic brain. Nat. Neurosci. 9, 340–348 199 Schmeisser, A., Christoph, M., Augstein, A., Marquetant, R., Kasper, M., Braun-Dullaeus, R. C. and Strasser, R. H. (2006) Apoptosis of human macrophages by Flt-4 signaling: implications for atherosclerotic plaque pathology. Cardiovasc. Res. 71, 774–784  c The Authors Journal compilation  c 2011 Biochemical Society

200 Petrova, T. V., Bono, P., Holnthoner, W., Chesnes, J., Pytowski, B., Sihto, H., Laakkonen, P., Heikkila, P., Joensuu, H. and Alitalo, K. (2008) VEGFR-3 expression is restricted to blood and lymphatic vessels in solid tumors. Cancer Cell 13, 554–556 201 Salameh, A., Galvagni, F., Bardelli, M., Bussolino, F. and Oliviero, S. (2005) Direct recruitment of CRK and GRB2 to VEGFR-3 induces proliferation, migration, and survival of endothelial cells through the activation of ERK, AKT, and JNK pathways. Blood 106, 3423–3431 202 Fournier, E., Dubreuil, P., Birnbaum, D. and Borg, J. P. (1995) Mutation at tyrosine residue 1337 abrogates ligand-dependent transforming capacity of the FLT4 receptor. Oncogene 11, 921–931 203 Galvagni, F., Pennacchini, S., Salameh, A., Rocchigiani, M., Neri, F., Orlandini, M., Petraglia, F., Gotta, S., Sardone, G. L., Matteucci, G. et al. (2010) Endothelial cell adhesion to the extracellular matrix induces c-Src-dependent VEGFR-3 phosphorylation without the activation of the receptor intrinsic kinase activity. Circ. Res. 106, 1839–1848 204 Alam, A., Herault, J. P., Barron, P., Favier, B., Fons, P., Delesque-Touchard, N., Senegas, I., Laboudie, P., Bonnin, J., Cassan, C. et al. (2004) Heterodimerization with vascular endothelial growth factor receptor-2 (VEGFR-2) is necessary for VEGFR-3 activity. Biochem. Biophys. Res. Commun. 324, 909–915 205 Goldman, J., Rutkowski, J. M., Shields, J. D., Pasquier, M. C., Cui, Y., Schmokel, H. G., Willey, S., Hicklin, D. J., Pytowski, B. and Swartz, M. A. (2007) Cooperative and redundant roles of VEGFR-2 and VEGFR-3 signaling in adult lymphangiogenesis. FASEB J. 21, 1003–1012 206 Ghalamkarpour, A., Morlot, S., Raas-Rothschild, A., Utkus, A., Mulliken, J. B., Boon, L. M. and Vikkula, M. (2006) Hereditary lymphedema type I associated with VEGFR3 mutation: the first de novo case and atypical presentations. Clin. Genet. 70, 330–335 207 Irrthum, A., Karkkainen, M. J., Devriendt, K., Alitalo, K. and Vikkula, M. (2000) Congenital hereditary lymphedema caused by a mutation that inactivates VEGFR3 tyrosine kinase. Am. J. Hum. Genet. 67, 295–301 208 Karkkainen, M. J., Ferrell, R. E., Lawrence, E. C., Kimak, M. A., Levinson, K. L., McTigue, M. A., Alitalo, K. and Finegold, D. N. (2000) Missense mutations interfere with VEGFR-3 signalling in primary lymphoedema. Nat. Genet. 25, 153–159 209 Karkkainen, M. J., Saaristo, A., Jussila, L., Karila, K. A., Lawrence, E. C., Pajusola, K., Bueler, H., Eichmann, A., Kauppinen, R., Kettunen, M. I. et al. (2001) A model for gene therapy of human hereditary lymphedema. Proc. Natl. Acad. Sci. U.S.A. 98, 12677–12682 210 Dumont, D. J., Jussila, L., Taipale, J., Lymboussaki, A., Mustonen, T., Pajusola, K., Breitman, M. and Alitalo, K. (1998) Cardiovascular failure in mouse embryos deficient in VEGF receptor-3. Science 282, 946–949 211 Haiko, P., Makinen, T., Keskitalo, S., Taipale, J., Karkkainen, M. J., Baldwin, M. E., Stacker, S. A., Achen, M. G. and Alitalo, K. (2008) Deletion of VEGF-C and VEGF-D is not equivalent to VEGFR-3-null in mouse embryos. Mol. Cell. Biol. 28, 4843–4850 212 Zhang, L., Zhou, F., Han, W., Shen, B., Luo, J., Shibuya, M. and He, Y. (2010) VEGFR-3 ligand-binding and kinase activity are required for lymphangiogenesis but not for angiogenesis. Cell Res. 20, 1319–1331 213 Tvorogov, D., Anisimov, A., Zheng, W., Leppanen, V. M., Tammela, T., Laurinavicius, S., Holnthoner, W., Helotera, H., Holopainen, T., Jeltsch, M. et al. (2010) Effective suppression of vascular network formation by combination of antibodies blocking VEGFR ligand binding and receptor dimerization. Cancer Cell 18, 630–640 214 Karkkainen, M. J., Haiko, P., Sainio, K., Partanen, J., Taipale, J., Petrova, T. V., Jeltsch, M., Jackson, D. G., Talikka, M., Rauvala, H. et al. (2004) Vascular endothelial growth factor C is required for sprouting of the first lymphatic vessels from embryonic veins. Nat. Immunol. 5, 74–80 215 Fruman, D. A., Mauvais-Jarvis, F., Pollard, D. A., Yballe, C. M., Brazil, D., Bronson, R. T., Kahn, C. R. and Cantley, L. C. (2000) Hypoglycaemia, liver necrosis and perinatal death in mice lacking all isoforms of phosphoinositide 3-kinase p85α. Nat. Genet. 26, 379–382 216 Zhou, F., Chang, Z., Zhang, L., Hong, Y. K., Shen, B., Wang, B., Zhang, F., Lu, G., Tvorogov, D., Alitalo, K. et al. (2010) Akt/Protein kinase B is required for lymphatic network formation, remodeling, and valve development. Am. J. Pathol. 177, 2124–2133 217 Ichise, T., Yoshida, N. and Ichise, H. (2010) H-, N- and Kras cooperatively regulate lymphatic vessel growth by modulating VEGFR3 expression in lymphatic endothelial cells in mice. Development 137, 1003–1013 218 Wang, Y., Nakayama, M., Pitulescu, M. E., Schmidt, T. S., Bochenek, M. L., Sakakibara, A., Adams, S., Davy, A., Deutsch, U., Luthi, U. et al. (2010) Ephrin-B2 controls VEGF-induced angiogenesis and lymphangiogenesis. Nature 465, 483–486 219 Kawamura, H., Li, X., Harper, S. J., Bates, D. O. and Claesson-Welsh, L. (2008) Vascular endothelial growth factor (VEGF)-A165b is a weak in vitro agonist for VEGF receptor-2 due to lack of coreceptor binding and deficient regulation of kinase activity. Cancer Res. 68, 4683–4692 220 Park, M. and Lee, S. T. (1999) The fourth immunoglobulin-like loop in the extracellular domain of FLT-1, a VEGF receptor, includes a major heparin-binding site. Biochem. Biophys. Res. Commun. 264, 730–734

Signal transduction by VEGF receptors 221 Xu, D., Fuster, M. M., Lawrence, R. and Esko, J. D. (2011) Heparan sulfate regulates VEGF165- and VEGF121-mediated vascular hyperpermeability. J. Biol. Chem. 286, 737–745 222 Vander Kooi, C. W., Jusino, M. A., Perman, B., Neau, D. B., Bellamy, H. D. and Leahy, D. J. (2007) Structural basis for ligand and heparin binding to neuropilin B domains. Proc. Natl. Acad. Sci. U.S.A. 104, 6152–6157 223 Stringer, S. E. (2006) The role of heparan sulphate proteoglycans in angiogenesis. Biochem. Soc. Trans. 34, 451–453 224 Jakobsson, L., Kreuger, J., Holmborn, K., Lundin, L., Eriksson, I., Kjellen, L. and Claesson-Welsh, L. (2006) Heparan sulfate in trans potentiates VEGFR-mediated angiogenesis. Dev. Cell 10, 625–634 225 Gerhardt, H. and Betsholtz, C. (2005) How do endothelial cells orientate? EXS 94, 3–15 226 Ruhrberg, C., Gerhardt, H., Golding, M., Watson, R., Ioannidou, S., Fujisawa, H., Betsholtz, C. and Shima, D. T. (2002) Spatially restricted patterning cues provided by heparin-binding VEGF-A control blood vessel branching morphogenesis. Genes Dev. 16, 2684–2698 227 Stalmans, I., Ng, Y. S., Rohan, R., Fruttiger, M., Bouche, A., Yuce, A., Fujisawa, H., Hermans, B., Shani, M., Jansen, S. et al. (2002) Arteriolar and venular patterning in retinas of mice selectively expressing VEGF isoforms. J. Clin. Invest. 109, 327–336 228 Fujisawa, H., Kitsukawa, T., Kawakami, A., Takagi, S., Shimizu, M. and Hirata, T. (1997) Roles of a neuronal cell-surface molecule, neuropilin, in nerve fiber fasciculation and guidance. Cell Tissue Res. 290, 465–470 229 Giger, R. J., Urquhart, E. R., Gillespie, S. K., Levengood, D. V., Ginty, D. D. and Kolodkin, A. L. (1998) Neuropilin-2 is a receptor for semaphorin IV: insight into the structural basis of receptor function and specificity. Neuron 21, 1079–1092 230 Soker, S., Takashima, S., Miao, H. Q., Neufeld, G. and Klagsbrun, M. (1998) Neuropilin-1 is expressed by endothelial and tumor cells as an isoform-specific receptor for vascular endothelial growth factor. Cell 92, 735–745 231 Chen, H., Chedotal, A., He, Z., Goodman, C. S. and Tessier-Lavigne, M. (1997) Neuropilin-2, a novel member of the neuropilin family, is a high affinity receptor for the semaphorins Sema E and Sema IV but not Sema III. Neuron 19, 547–559 232 He, Z. and Tessier-Lavigne, M. (1997) Neuropilin is a receptor for the axonal chemorepellent semaphorin III. Cell 90, 739–751 233 Kolodkin, A. L., Levengood, D. V., Rowe, E. G., Tai, Y. T., Giger, R. J. and Ginty, D. D. (1997) Neuropilin is a semaphorin III receptor. Cell 90, 753–762 234 Vander Kool, C. W., Jusino, M. A., Perman, B., Neau, D. B., Bellamy, H. D. and Leahy, D. J. (2007) Structural basis for ligand and heparin binding to neuropilin B domains. Proc. Natl. Acad. Sci. U.S.A. 104, 6152–6157 235 Pan, Q., Chathery, Y., Wu, Y., Rathore, N., Tong, R. K., Peale, F., Bagri, A., Tessier-Lavigne, M., Koch, A. W. and Watts, R. J. (2007) Neuropilin-1 binds to VEGF121 and regulates endothelial cell migration and sprouting. J. Biol. Chem. 282, 24049–24056 236 Karpanen, T., Heckman, C. A., Keskitalo, S., Jeltsch, M., Ollila, H., Neufeld, G., Tamagnone, L. and Alitalo, K. (2006) Functional interaction of VEGF-C and VEGF-D with neuropilin receptors. FASEB J. 20, 1462–1472 237 Geretti, E., Shimizu, A., Kurschat, P. and Klagsbrun, M. (2007) Site-directed mutagenesis in the B-neuropilin-2 domain selectively enhances its affinity to VEGF165, but not to semaphorin 3F. J. Biol. Chem. 282, 25698–25707 238 Shintani, Y., Takashima, S., Asano, Y., Kato, H., Liao, Y., Yamazaki, S., Tsukamoto, O., Seguchi, O., Yamamoto, H., Fukushima, T. et al. (2006) Glycosaminoglycan modification of neuropilin-1 modulates VEGFR2 signaling. EMBO J. 25, 3045–3055 239 Favier, B., Alam, A., Barron, P., Bonnin, J., Laboudie, P., Fons, P., Mandron, M., Herault, J. P., Neufeld, G., Savi, P. et al. (2006) Neuropilin-2 interacts with VEGFR-2 and VEGFR-3 and promotes human endothelial cell survival and migration. Blood 108, 1243–1250 240 Wang, L., Zeng, H., Wang, P., Soker, S. and Mukhopadhyay, D. (2003) Neuropilin-1-mediated vascular permeability factor/vascular endothelial growth factor-dependent endothelial cell migration. J. Biol. Chem. 278, 48848–48860 241 Becker, P. M., Waltenberger, J., Yachechko, R., Mirzapoiazova, T., Sham, J. S., Lee, C. G., Elias, J. A. and Verin, A. D. (2005) Neuropilin-1 regulates vascular endothelial growth factor-mediated endothelial permeability. Circ. Res. 96, 1257–1265 242 Kawamura, H., Li, X., Goishi, K., van Meeteren, L. A., Jakobsson, L., Cebe-Suarez, S., Shimizu, A., Edholm, D., Ballmer-Hofer, K., Kjellen, L. et al. (2008) Neuropilin-1 in regulation of VEGF-induced activation of p38MAPK and endothelial cell organization. Blood 112, 3638–3649 243 Cai, H. and Reed, R. R. (1999) Cloning and characterization of neuropilin-1-interacting protein: a PSD-95/Dlg/ZO-1 domain-containing protein that interacts with the cytoplasmic domain of neuropilin-1. J. Neurosci. 19, 6519–6527

183

244 Frankel, P., Pellet-Many, C., Lehtolainen, P., D’Abaco, G. M., Tickner, M. L., Cheng, L. and Zachary, I. C. (2008) Chondroitin sulphate-modified neuropilin 1 is expressed in human tumour cells and modulates 3D invasion in the U87MG human glioblastoma cell line through a p130Cas-mediated pathway. EMBO Rep. 9, 983–989 245 Eichmann, A., Makinen, T. and Alitalo, K. (2005) Neural guidance molecules regulate vascular remodeling and vessel navigation. Genes Dev. 19, 1013–1021 246 Herzog, Y., Kalcheim, C., Kahane, N., Reshef, R. and Neufeld, G. (2001) Differential expression of neuropilin-1 and neuropilin-2 in arteries and veins. Mech. Dev. 109, 115–119 247 Yuan, L., Moyon, D., Pardanaud, L., Breant, C., Karkkainen, M. J., Alitalo, K. and Eichmann, A. (2002) Abnormal lymphatic vessel development in neuropilin 2 mutant mice. Development 129, 4797–4806 248 Fujisawa, H. and Kitsukawa, T. (1998) Receptors for collapsin/semaphorins. Curr. Opin. Neurobiol. 8, 587–592 249 Xu, Y., Yuan, L., Mak, J., Pardanaud, L., Caunt, M., Kasman, I., Larrivee, B., Del Toro, R., Suchting, S., Medvinsky, A. et al. (2010) Neuropilin-2 mediates VEGF-C-induced lymphatic sprouting together with VEGFR3. J. Cell Biol. 188, 115–130 250 Takashima, S., Kitakaze, M., Asakura, M., Asanuma, H., Sanada, S., Tashiro, F., Niwa, H., Miyazaki Ji, J., Hirota, S., Kitamura, Y. et al. (2002) Targeting of both mouse neuropilin-1 and neuropilin-2 genes severely impairs developmental yolk sac and embryonic angiogenesis. Proc. Natl. Acad. Sci. U.S.A. 99, 3657–3662 251 Pan, Q., Chanthery, Y., Liang, W. C., Stawicki, S., Mak, J., Rathore, N., Tong, R. K., Kowalski, J., Yee, S. F., Pacheco, G. et al. (2007) Blocking neuropilin-1 function has an additive effect with anti-VEGF to inhibit tumor growth. Cancer Cell 11, 53–67 252 Caunt, M., Mak, J., Liang, W. C., Stawicki, S., Pan, Q., Tong, R. K., Kowalski, J., Ho, C., Reslan, H. B., Ross, J. et al. (2008) Blocking neuropilin-2 function inhibits tumor cell metastasis. Cancer Cell 13, 331–342 253 Cackowski, F. C., Xu, L., Hu, B. and Cheng, S. Y. (2004) Identification of two novel alternatively spliced neuropilin-1 isoforms. Genomics 84, 82–94 254 Borges, E., Jan, Y. and Ruoslahti, E. (2000) Platelet-derived growth factor receptor β and vascular endothelial growth factor receptor 2 bind to the β3 integrin through its extracellular domain. J. Biol. Chem. 275, 39867–39873 255 Liu, Z., Kobayashi, K., van Dinther, M., van Heiningen, S. H., Valdimarsdottir, G., van Laar, T., Scharpfenecker, M., Lowik, C. W., Goumans, M. J., Ten Dijke, P. and Pardali, E. (2009) VEGF and inhibitors of TGFβ type-I receptor kinase synergistically promote blood-vessel formation by inducing α5-integrin expression. J. Cell Sci. 122, 3294–3302 256 Mahabeleshwar, G. H., Chen, J., Feng, W., Somanath, P. R., Razorenova, O. V. and Byzova, T. V. (2008) Integrin affinity modulation in angiogenesis. Cell Cycle 7, 335–347 257 Pampori, N., Hato, T., Stupack, D. G., Aidoudi, S., Cheresh, D. A., Nemerow, G. R. and Shattil, S. J. (1999) Mechanisms and consequences of affinity modulation of integrin α V β 3 detected with a novel patch-engineered monovalent ligand. J. Biol. Chem. 274, 21609–21616 258 Robinson, S. D., Reynolds, L. E., Kostourou, V., Reynolds, A. R., da Silva, R. G., Tavora, B., Baker, M., Marshall, J. F. and Hodivala-Dilke, K. M. (2009) αVβ3 integrin limits the contribution of neuropilin-1 to vascular endothelial growth factor-induced angiogenesis. J. Biol. Chem. 284, 33966–33981 259 Mahabeleshwar, G. H., Feng, W., Phillips, D. R. and Byzova, T. V. (2006) Integrin signaling is critical for pathological angiogenesis. J. Exp. Med. 203, 2495–2507 260 Miao, H. Q., Soker, S., Feiner, L., Alonso, J. L., Raper, J. A. and Klagsbrun, M. (1999) Neuropilin-1 mediates collapsin-1/semaphorin III inhibition of endothelial cell motility: functional competition of collapsin-1 and vascular endothelial growth factor-165. J. Cell Biol. 146, 233–242 261 Zhang, X., Groopman, J. E. and Wang, J. F. (2005) Extracellular matrix regulates endothelial functions through interaction of VEGFR-3 and integrin α5β1. J. Cell. Physiol. 202, 205–214 262 Bergers, G. and Hanahan, D. (2008) Modes of resistance to anti-angiogenic therapy. Nat. Rev. Cancer. 8, 592–603 263 Ebos, J. M., Lee, C. R., Cruz-Munoz, W., Bjarnason, G. A., Christensen, J. G. and Kerbel, R. S. (2009) Accelerated metastasis after short-term treatment with a potent inhibitor of tumor angiogenesis. Cancer Cell 15, 232–239 264 Paez-Ribes, M., Allen, E., Hudock, J., Takeda, T., Okuyama, H., Vinals, F., Inoue, M., Bergers, G., Hanahan, D. and Casanovas, O. (2009) Antiangiogenic therapy elicits malignant progression of tumors to increased local invasion and distant metastasis. Cancer Cell 15, 220–231 265 Tugues, S., Koch, S., Gualandi, L., Li, X. and Claesson-Welsh, L. (2011) Vascular endothelial growth factors and receptors; anti-angiogenic therapy in treatment of cancer. Mol. Aspects Med., doi:10.1016/j.mam.2011.04.004

Received 15 February 2011/12 April 2011; accepted 15 April 2011 Published on the Internet 28 June 2011, doi:10.1042/BJ20110301  c The Authors Journal compilation  c 2011 Biochemical Society