Skip to main content
Erschienen in: Angiogenesis 3/2021

Open Access 13.05.2021 | Original Paper

The endosomal RIN2/Rab5C machinery prevents VEGFR2 degradation to control gene expression and tip cell identity during angiogenesis

verfasst von: Lanette Kempers, Yuki Wakayama, Ivo van der Bijl, Charita Furumaya, Iris M. De Cuyper, Aldo Jongejan, Marije Kat, Anne-Marieke D. van Stalborch, Antonius L. van Boxtel, Marvin Hubert, Dirk Geerts, Jaap D. van Buul, Dirk de Korte, Wiebke Herzog, Coert Margadant

Erschienen in: Angiogenesis | Ausgabe 3/2021

Abstract

Sprouting angiogenesis is key to many pathophysiological conditions, and is strongly regulated by vascular endothelial growth factor (VEGF) signaling through VEGF receptor 2 (VEGFR2). Here we report that the early endosomal GTPase Rab5C and its activator RIN2 prevent lysosomal routing and degradation of VEGF-bound, internalized VEGFR2 in human endothelial cells. Stabilization of endosomal VEGFR2 levels by RIN2/Rab5C is crucial for VEGF signaling through the ERK and PI3-K pathways, the expression of immediate VEGF target genes, as well as specification of angiogenic ‘tip’ and ‘stalk’ cell phenotypes and cell sprouting. Using overexpression of Rab mutants, knockdown and CRISPR/Cas9-mediated gene editing, and live-cell imaging in zebrafish, we further show that endosomal stabilization of VEGFR2 levels is required for developmental angiogenesis in vivo. In contrast, the premature degradation of internalized VEGFR2 disrupts VEGF signaling, gene expression, and tip cell formation and migration. Thus, an endosomal feedforward mechanism maintains receptor signaling by preventing lysosomal degradation, which is directly linked to the induction of target genes and cell fate in collectively migrating cells during morphogenesis.
Hinweise

Supplementary Information

The online version of this article contains supplementary material available (https://​doi.​org/​10.​1007/​s10456-021-09788-4).
Lanette Kempers, Yuki Wakayama, and Ivo van der Bijl have contributed equally to this work.
Wiebke Herzog and Coert Margadant are senior co-authors

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Abkürzungen
ALS2
Alsin
ALS2CL
ALS2 C-terminal-like protein
ANGPT2
Angiopoietin-2
APLN
Apelin
CA
Constitutively active
Ctrl
Control
DLAV
Dorsal longitudinal anastomotic vessel
DLL4
Delta Like ligand-4
DN
Dominant-Negative
EE
Early Endosome
EEA-1
Early Endosomal Autoantigen-1
ERK
Extracellular-signal Regulated Kinase
FDR
False discovery rate
GAPVD1
GTPase Activating Protein And VPS9 Domains 1
GEF
Guanine Exchange Factor
GEO
Gene Expression Omnibus
GFP
Green fluorescent protein
HUVEC
Human Umbilical Vein Endothelial Cells
hpf
Hours post-fertilization
ISV
Intersegmental vessel
MO
Morpholino
NID2
Nidogen-2
NR4A
Nuclear Receptor Subfamily 4 Group A
NRP2
Neuropilin-2
ORF
Open reading frame
PDGFB
Platelet-Derived Growth Factor B
PI3-K
Phosphatidylinositol 3-kinase
qPCR
Quantitative PCR
RABEP-2
Rab GTPase-binding effector protein 2
RIN
Ras and Rab Interactor
RINL
Ras and Rab INteractor-Like
sh
Short hairpin
TGN
Trans-Golgi network
VE-cadherin
Vascular Endothelial-cadherin
VEGF
Vascular Endothelial Growth Factor
VEGFR
Vascular Endothelial Growth Factor Receptor
WT
Wild-type

Introduction

Sprouting angiogenesis is crucial for a range of pathophysiological processes including embryonic development, wound healing, tissue remodeling, cancer, and cardiovascular disease [1]. A key pro-angiogenic event is the interaction of vascular endothelial growth factor A (VEGF-A, further referred to as VEGF) with VEGF receptor 2 (VEGFR2), which promotes endothelial proliferation, survival, and migration through activation of the extracellular-signal regulated kinase (ERK) and phosphatidylinositol 3-kinase (PI3-K) pathways [2]. VEGF/VEGFR2 signaling also induces a pro-angiogenic gene expression program and stimulates the selection of specialized ‘tip cells’, which have a distinct morphology and transcriptional signature, and promote the guidance of nascent sprouts [37]. Tip cells are followed by ‘stalk cells’, which promote sprout elongation and lumenization [47]. VEGF-dependent specification of tip and stalk cells is fine-tuned by activation of Notch signaling, which downregulates VEGFR2 expression to prevent uncontrolled sprouting [710], as well as by the decoy receptor VEGFR1, which sequesters VEGF away from VEGFR2 [1113].
Upon VEGFR2 activation by ligand binding, VEGFR2 is rapidly internalized and subsequently recruited to early endosomes (EEs) [2, 1418], which are marked by the presence of GTPases of the Rab5 subfamily [19]. Rab GTPases are key regulators of vesicle fusion and protein sorting that are activated by GDP/GTP exchange, catalyzed by guanine nucleotide exchange factors (GEFs) [1921]. Activation is crucial for Rab localization to intracellular membranes, and for the recruitment of effector proteins [1921]. From EEs, VEGFR2 is trafficked to lysosomes and degraded, thus accelerating its turnover and terminating VEGF signaling, in a manner dependent on epsin binding to ubiquitinated cytoplasmic motifs in VEGFR2 [16, 2225]. Alternatively, VEGFR2 is recycled back to the cell-surface in recycling compartments, containing either Rab4 or Rab11 [14, 16, 17].
Internalization of activated VEGFR2 and its subsequent flux through the endosomal system are regulated by endocytic adapters, as well as by cell-surface receptors that associate with VEGFR2 [14, 23, 2531]. Although it is known that VEGFR2 can signal from endosomal compartments [2, 18], the interrelation between VEGFR2 endocytosis and signaling is complex. For example, ephrin B2 knockout impairs both VEGFR2 endocytosis and signaling [30, 31], whereas the targeted deletion of epsins or dynamin blocks VEGFR2 internalization but not signaling [28, 32]. Furthermore, it is important that VEGFR2-containing endosomes progress normally through the endosomal system, as delayed trafficking exposes internalized VEGFR2 to selective dephosphorylation by tyrosine phosphatase PTP1b, leading to reduced activation of the ERK, but not PI3-K pathway, and impaired arteriogenesis [26, 33, 34].
While these findings emphasize the importance of endocytosis for VEGF signaling, less is known about the mechanisms and factors that maintain endosomal VEGFR2 levels and prevent premature degradation. Studies in zebrafish have shown that Rab5C is particularly highly expressed in endothelium and important for endothelial Notch trafficking [3538], but its role in VEGFR2 traffic is unknown. Moreover, it is unclear if and how endosomal VEGF signaling regulates the induction of VEGF target genes, or the acquisition of tip versus stalk cell properties during sprouting angiogenesis.
Here we show that the regulated maintenance of internalized VEGFR2 and its diversion from the degradation pathway is tightly linked to the induction of VEGF target genes, and is crucial for tip cell specification and endothelial cell migration. The endosomal VEGFR2 pool is protected from degradation by Rab5C, which is recruited to endosomes by the GEF RIN2. Finally, we show by a number of in vitro and in vivo approaches that manipulation of the RIN2/Rab5C machinery leads to premature VEGFR2 degradation, thus disturbing normal endothelial VEGF/Notch signaling and VEGF-dependent gene expression, generation of functional tip cells, and sprouting angiogenesis. In summary, an endosomal feedforward loop controlled by RIN2/Rab5C prevents VEGFR2 degradation, and maintains a VEGF signaling window required for VEGF-induced gene expression, tip/stalk cell specification, and vascular sprouting.

Results

Rab5C protects VEGFR2 from VEGF-induced lysosomal degradation

We first depleted Rab5C from human umbilical vein endothelial cells (HUVECs) with shRNA pools (leading to ~ 90% reduction at the mRNA level; Fig S1A), and assessed the cell-surface levels of VEGFR2 by flow cytometry. Intriguingly, VEGFR2 surface levels in steady-state HUVECs (growing in the presence of VEGF) were consistently reduced two-fold in sh_Rab5C cells, compared to HUVECs expressing scrambled sequences as a control (sh_Ctrl) (Fig. 1a). The decrease in VEGFR2 was further confirmed by quantification of the total VEGFR2 protein levels by Western blotting (Fig. 1b). In contrast, VEGFR1 protein levels were not affected, suggesting that VEGFR1 is not regulated by Rab5C (Fig. 1b; Fig S1B). We then investigated if and how Rab5C regulates VEGFR2 endocytic recycling using flow cytometry. To test whether the internal pool of VEGFR2 can be recruited normally to the plasma membrane, we starved HUVECs for 30 min, a time-frame that allows mobilization of the internal VEGFR2 pool to the plasma membrane but is too short to induce strong changes in de novo protein synthesis. Starvation significantly increased the cell-surface levels of VEGFR2 compared to those in steady-state in sh_Ctrl but not sh_Rab5C cells, indicating that the recruitment from intracellular compartments was impaired (Fig. 1c, d). VEGF induced VEGFR2 endocytosis in both sh_Ctrl and sh_Rab5C cells, although in the latter the internalization was slightly reduced (Fig. 1c, e). The endosomal pool of VEGFR2 is protected from degradation in the absence of ligand, while VEGF stimulation results in rapid degradation [16, 22]. Therefore, we hypothesized that the reduced levels of VEGFR2 in Rab5C-depleted cells were due to increased degradation. To assess VEGFR2 degradation, HUVECs were deprived of growth factors overnight and then stimulated with VEGF, which resulted in a rapid decrease of up to 40% of VEGFR2 levels in 30 min (Fig. 1f), consistent with previous reports [16, 22]. Strikingly, VEGFR2 degradation was enhanced twofold in Rab5C-depleted cells, confirming our hypothesis that the reduction in VEGFR2 levels in these cells is caused by increased degradation (Fig. 1f). Moreover, inhibition of lysosomal proteases blocked VEGF-induced degradation, and nullified the differences in VEGFR2 degradation between sh_Ctrl and sh_Rab5C cells (Fig S1C). In HUVECs that were maintained in the absence of VEGF, VEGFR2 levels declined only modestly (20% over 120 min), and this was not significantly altered by depletion of Rab5C (Fig. 1g).
In summary, these results show that Rab5C protects the endosomal VEGFR2 pool from VEGF-induced lysosomal degradation.

Rab5C regulates VEGF signaling, the VEGF-induced immediate transcriptome, and endothelial cell sprouting

We next investigated if and how Rab5C regulates VEGF signaling. For this purpose, sh_Ctrl and sh_Rab5C HUVECs were starved overnight and then stimulated with VEGF, whereafter AKT phosphorylation was assessed as a read-out for PI3-K-mediated VEGF signaling. VEGF triggered robust and persistent AKT phosphorylation within 10 min in sh_Ctrl cells, which was almost completely abolished in Rab5C-depleted cells, indicating that Rab5C is required for VEGF-induced PI3-K signaling (Fig. 2a). We also investigated VEGF-induced ERK-1/2 phosphorylation. In sh_Ctrl cells, robust induction of ERK-1/2 phosphorylation was observed within 5 min after VEGF addition and persisted for up to 30 min, while this was significantly reduced in sh_Rab5C cells (Fig. 2b).
We then assessed if and how the reduced VEGFR2 signaling observed upon depletion of Rab5C impacts on VEGF-regulated gene expression by mRNA sequencing analysis (RNA-seq). Differential expression was assessed by empirical Bayes analysis followed by correction of p values for multiple testing using the Benjamini–Hochberg false discovery rate (FDR), with a cut-off adjusted p value of 0.05. Using these criteria, the expression of 40 genes was significantly induced and 1 gene was repressed in sh_Ctrl cells after 1 h (t = 1) of VEGF stimulation, while after 4 h (t = 4) the expression of these genes was largely back to base-line levels (Table S1 and Fig. 2c). These kinetics are comparable with those found in previous reports [3941]. The induced target genes include well-known immediate VEGF targets, 58% of which encoding transcriptional regulators such as Krüppel-like factors, nuclear receptors of the NR4A family, and members of the Fos/Jun and early growth response protein families, also consistent with earlier findings (Table S1 and Fig. 2c) [3941]. Strikingly, expression of a large number of VEGF target genes (34 out of 41) was altered in sh_Rab5C cells (Fig. 2c), and geneset enrichment analysis revealed that at t = 1, expression of the ‘VEGF transcriptome’ was significantly different with respect to that in sh_Ctrl cells (p = 3.12*10–4, FDR = 1.52*10–2).
We then assessed the role of Rab5C in sprouting angiogenesis, using the fibrin bead assay [42, 43]. Quantification of the extent of sprouting after 48 h revealed that depletion of Rab5C strongly impaired sprouting, by affecting both the numbers of formed sprouts and their length (Fig. 2d–g). In addition, tip cells clearly formed multiple filopodia in the sh_Ctrl population, a process stimulated by VEGF signaling, whereas sh_Rab5C tip cells were on average more ‘blunted’ (Fig. 2d).
Together, these data show that Rab5C promotes endosomal VEGFR2 signaling and full induction of the immediate VEGF-induced transcriptome, and that Rab5C is required for efficient sprouting angiogenesis.

Rab5C promotes tip cell formation

Because VEGF signaling and the expression of certain VEGF targets is important for the induction and maintenance of tip cells during sprouting angiogenesis, we next addressed whether the reduced sprouting observed in the absence of Rab5C is due to differences in VEGF-induced tip cell specification. For this purpose, we first investigated the expression of a panel of genes commonly associated with tip cell identity, including ANGPT2, APLN, DLL4, NID2, NRP2, PDGFB, TIE1, UNC5B, VEGFR2, and VEGFR3 [6, 7]. Intriguingly, depletion of Rab5C significantly suppressed the transcription of ANGPT2, APLN, DLL4, NID2, UNC5B, and VEGFR3, while the expression of NRP2, TIE1, and VEGFR2 was mildly but not significantly decreased (Fig. 3a). Of note, PDGFB expression appeared not to be regulated at all by Rab5C.
We then assessed tip cell formation in mosaic sprouting assays, by differentially labeling sh_Ctrl and sh_Rab5C cells using two distinct CellTracker dyes, mixing them in a 1:1 ratio, and determining the fraction of tip cells formed by each population after 48 h (Fig. 3b). The knockdown of Rab5C significantly reduced the numbers of tip cells in this assay (Fig. 3b). Similar results were obtained using a pool of mixed shRNAs (Fig. 3b), two different individual shRNAs (Fig S2A–C), and when the dyes were swapped between conditions (Fig S2B).
To address the role of Rab5C in an in vivo model system for sprouting angiogenesis, we next analyzed if Rab5C regulates intersegmental vessel (ISV) development in zebrafish. The zebrafish rab5c gene is highly homologous to its human counterpart, suggesting an important function in the vasculature [35, 37, 38]. We first generated a dominant-negative (DN) mutant Rab5C protein, carrying a single substitution (S35N) in the GTP-binding pocket (Fig S3A). DN Rab mutants sequester GEFs but fail to bind GTP, and therefore cannot be activated [44]. The mutant was fused to mCherry and expressed in HUVECs to test its ability to localize on endosomes. Wild-type (WT) Rab5C localized predominantly on endosomes, some of which were positive for EEA-1, while occasional distribution to the Golgi network was also observed, using Trans-Golgi Network (TGN)46 as a marker (Fig S3B). In contrast, the mutant localized predominantly at the Golgi and hardly on endosomes (Fig S3B). We then cloned the mCherry-tagged WT or DN human RAB5C gene into a construct that integrates via Tol2-mediated transgenesis and is driven by the zebrafish fli-1a promoter, to achieve expression specifically in vessels (Fig. 3c) [45]. The construct was injected into single-cell stage Tg(kdrl:GFP)s843 zebrafish embryos expressing GFP in all vascular endothelial cells, and mCherry-positive endothelial cells were examined in developing ISVs at 30–32 h post-fertilization (hpf) (Fig. 3d, e) [46, 47]. As these injections result in mosaic expression of mCherry-Rab5C in endothelial cells, mCherry-positive endothelial cells were scored for their position within the developing vessels (tip, stalk, or dorsal aorta; Fig S4A). The distribution of WT-Rab5C positive cells was similar to that of mCherry-expressing cells (Fig. 3f and Fig S4B). However, cells expressing DN-Rab5C were less commonly found in the tip cell position (22% versus 56% in the mCherry control), and most DN-Rab5C positive cells localized to the stalk or the dorsal aorta (Fig. 3f and Fig S4B).
Altogether, our findings indicate that Rab5C is required for tip cell formation, with Rab5C expressing cells having a competitive advantage for tip cell positioning over cells with impaired Rab5C function.

Rab5C is required for maintenance of Vegfr2 levels, Vegf/Notch signaling, and sprouting angiogenesis in zebrafish

We showed by mosaic overexpression of DN-Rab5C that Rab5C activation is required for tip cell formation in vivo. While these experiments allowed us to identify the preference for tip or stalk cell position in a competitive situation, we also analyzed the effects of DN-Rab5C overexpressed simultaneously in all endothelial cells, by generating transgenic Tg(kdrl:GFP)s843 embryos with vascular-specific expression of human mCherry-WT-Rab5C (Tg(fli1a:mCherry-WT-hRAB5C)mu227, or mCherry-DN-Rab5C (Tg(fli1a:mCherry-DN-hRAB5C)mu228. Similar to our findings in HUVECs, we observed localization of WT Rab5C protein in rapidly moving vesicles in endothelial cells in vivo (Fig. S4C), while the DN mutant was mostly absent from these vesicles but was instead retained in a perinuclear compartment, most likely the Golgi (Fig. S4C). Indeed, quantification confirmed that the number of mCherry-positive vesicles in DN-Rab5C-expressing ISV cells was strongly decreased, compared to WT-Rab5C expressing cells (Fig. S4D).
We then scored ISV formation at 30–32 hpf, a time-point at which normal ISV endothelial cells have migrated dorsally and interconnected, to form the dorsal longitudinal anastomotic vessel (DLAV) [46, 47]. The survival of DN-Rab5C-expressing embryos was compromised at later stages but not before 30–32 hpf, thus allowing analysis of the developing ISVs. Without WT cells taking over the tip cell function, we observed an impairment of endothelial cell migration in the ISVs of embryos expressing DN-Rab5C (Fig. 4a). While in WT-Rab5C-expressing embryos ISV development was characterized by 83.3% reaching the DLAV and 16.7% migrating nearly to the top, DN-Rab5C-overexpressing endothelial cells migrated only to the level of the horizontal myoseptum (‘half’) in 31% of the embryos, with 47.3% reaching the ISV length, and only 21.7% reaching the DLAV (Fig. 4b).
To disrupt zebrafish Rab5c function in vivo via additional methods, we also used Morpholino (MO)-mediated block of rab5c translation (rab5c ATG MO) or splicing (rab5c e2i2 MO), as well as CRISPR/Cas9-mediated genome modification to induce mutations in the rab5c gene (Fig. S5A–E). A heterozygous rab5c mutant line was generated by injections of rab5c specific guide RNAs (gRNAs) and Cas9 protein, growth to adulthood, identification of germline-transmitting adults, growth of the F1 generation, and identification of the mutation. Detailed genomic analysis revealed that the resulting rab5c mutant (designated rab5cmu229) contained a premature stop codon, thus terminating protein synthesis at 29 (instead of 221) amino acids (Fig. S5D,E). We then analyzed vascular development in MO-injected embryos and in the homozygous mutant offspring of rab5c+/mu229 parents. Similar to the DN-Rab5C-expressing embryos, the homozygous mutants died during later stages of development. We observed mispatterning of the ISVs and impaired dorsal migration in individual ISV sprouts in embryos injected with rab5c MOs (compared to Control MOs), as well as in rab5cmu229/mu229 embryos (compared to rab5c+/+ siblings) (Fig. 4c,d), thus confirming that Rab5c is required for tip cell formation and normal ISV migration in vivo. Additionally, abnormal sprouting angiogenesis was also revealed by live imaging of ISV development using time-lapse microscopy (Movies S1,S2 and Fig. S6A).
We also investigated the levels of zebrafish Vegfr2 (Kdrl) protein in embryos injected with Control MO or rab5c MO by Western blotting (Fig. 5a). Similar to what we observed in HUVECs, a significant reduction in Kdrl protein levels was apparent in Rab5c-depleted embryos (Fig. 5a,b). Consistently, tip cells in these embryos formed shorter filopodia, as revealed by Rab5c knockdown in Tg(fli1a:lifeact-EGFP)mu240 embryos, which express LifeAct-eGFP in the vasculature to visualize the actin cytoskeleton (Movies S3,S4 and Fig. 5c) [48]. Vegf-induced tip cell specification in zebrafish requires initial activation of Notch signaling in tip cells, which subsequently directs them into developing arteries [49, 50]. Since we found an impairment of tip cell morphology and behavior in Rab5c-depleted embryos, we also addressed whether Notch signaling was affected. For this purpose we used the Tg(TP1bglob:VenusPEST)s940 transgenic line to image the highly dynamic expression of a very short-lived Venus protein (Venus-PEST), that is driven by the TP1 promoter element and thus reports activation of Notch signaling [51]. Notch activation was visible in arteries, as well as in some tip cells and arterial ISVs (Fig. 5d), in addition to a number of non-vascular mesenchymal cells. As expected from our morphological data, a strong reduction of Venus-PEST expression in rab5c MO-injected embryos was observed, indicating that activation of Notch signaling was reduced and further supporting our hypothesis that Rab5c is required for efficient generation of functional tip cells (Fig. 5d).
Altogether, these data show that in line with our observations in HUVECs (Figs. 13), Rab5c is crucial for the maintenance of Vegfr2 levels in zebrafish, and thereby regulates tip cell formation and sprouting angiogenesis in vivo.

The Rab5 GEF RIN2 maintains VEGFR2 levels in endothelial cells and promotes angiogenic sprouting

The previous sections have shown that activation of endothelial Rab5C is required for sprouting angiogenesis. Because Rabs are activated by GEFs, we next investigated which of the eight known GEFs for Rab5 GTPases are expressed in endothelial cells, using publicly available genome-wide mRNA expression profiles. This analysis suggested that HUVECs express at least seven different Rab5 GEFs, while RINL was not represented on the used arrays (Fig. 6a and Table S2). To investigate which of these GEFs is involved in the Rab5C-dependent effects on VEGFR2 and angiogenic sprouting, we silenced the expression of each individual GEF, including RINL, and assessed VEGFR2 cell-surface levels by flow cytometry. Of all investigated GEFs, only the depletion of Ras and Rab interactor (RIN) 2 significantly reduced the surface levels of VEGFR2, to the same extent as the depletion of Rab5C (Fig. 6b). Furthermore, total protein levels of VEGFR2 were also reduced by RIN2 depletion, as assessed by Western blotting (Fig. 6c, d). As in the case of Rab5C depletion, the reduction of VEGFR2 levels was associated with an increase in VEGF-induced VEGFR2 degradation (Fig. 6e). Finally, we performed a sprouting assay using RIN2-depleted cells. As expected, both the number of formed sprouts as well as the total network length were suppressed by RIN2 depletion (Fig. 6f–i).
Collectively, our data show that knockdown of RIN2 expression recapitulates the effects of Rab5C depletion, and strongly suggest that RIN2-mediated Rab5C activation prevents VEGFR2 degradation to promote sprouting angiogenesis.

RIN2 regulates Rab5C recruitment and is required for sprouting angiogenesis in vivo

We next tested whether forced Rab5C activation can circumvent the requirement for RIN2 to drive sprouting. For this purpose, we generated a constitutively active (CA) mutant (Q80L) of human Rab5C, which blocks GTP hydrolysis, fused to mScarlet (Fig. S6B). Expression of this mutant into HUVECs caused enlargement of EEs, indicative of constitutive Rab5C activation (Fig. S6C) [52]. We then expressed mScarlet-CA-Rab5C into RIN2-depleted cells, which slightly lowered endogenous Rab5C expression, whereas RIN2 knockdown by itself did not affect endogenous Rab5C levels or vice versa (Fig. 7a; Fig. S6D). Importantly, expression of CA-Rab5C in RIN2-depleted HUVECs rescued the sprouting defects caused by RIN2 deficiency, and thus bypassed the requirement for RIN2 (Fig. 7b). These data confirm that 1) Rab5C activation promotes sprouting angiogenesis, and 2) the primary role of RIN2 in sprouting angiogenesis is to activate Rab5C.
To assess the effects of RIN2 knockdown on Rab5C function in vivo, we generated a splice-blocking MO (e3i3) against the zebrafish rin2 gene (Fig. S7A,B), which was injected into Tg(fli1a:mCherry-WT-hRAB5C)mu227;Tg(kdrl:GFP)s843 zebrafish embryos to examine Rab5C localization. Depletion of zebrafish Rin2 protein resulted in a partial failure of Rab5C to localize in vesicular compartments (Fig. 7c), indicating that the recruitment of Rab5C requires Rin2 and hence, that the Rin2-Rab5c interaction is functionally conserved in zebrafish.
We also addressed the functional role of zebrafish Rin2 in sprouting angiogenesis in Tg(kdrl:GFP)s843 embryos. Knockdown of Rin2 resulted in impaired endothelial cell migration in the ISVs, similar to that induced by deficiency of Rab5c (Fig. 7d, e). The migration defect was again accompanied by reduced activation of Notch signaling, as analyzed by Tp1:Venus-Pest expression (Fig. S7C).
In summary, these results indicate that Rin2 is crucial for Rab5C activation and recruitment in endothelial cells, which regulates Vegf/Notch signaling, gene expression, tip/stalk cell specification, and sprouting angiogenesis (Fig. 7f).

Discussion

In this study, we identify an endosomal regulatory mechanism dependent on Rab5C and RIN2 that is crucial for receptor signaling, transcriptional output, and sprouting angiogenesis. Our data indicate that RIN2 recruits Rab5C to EEs, where it prevents the degradation of internalized VEGFR2. This is required to maintain VEGFR2 levels and to sustain endosomal VEGF/VEGFR2 signaling toward PI3-K and ERK pathways. Furthermore, it is necessary for the normal expression of VEGF target genes, including various tip cell markers, as well as for the generation of functional tip cells and their migration (summarized in Fig. 7f). Our data emphasize the crucial role of endosomal trafficking for cell fate decisions, and have important implications.
First, the amplitude and duration of VEGF signaling toward PI3-K and to a lesser extent ERK is dependent on endosomal Rab5C. This observation implies that VEGF-induced PI3-K signaling occurs primarily from EEs and much less from the plasma membrane, which fits with the well-documented role of PI3-K as a direct effector of Rab5 GTPases on EEs [19]. Although it is possible that Rab5C depletion indirectly affects ERK signaling from cell-surface VEGFR2 due to the decrease of VEGFR2 levels, it seems likely that also VEGF-induced ERK signaling occurs mainly from EEs, which is consistent with the findings of others and the existence of an ERK scaffold complex on endosomes [2, 34, 53]. While VEGFR2 signaling in the absence of internalization can also occur from the plasma membrane [28, 32], our data suggest that once VEGFR2 is endocytosed, components of the endosomal machinery, such as Rab5C, are required for normal VEGF signaling.
Second, our findings show that endosomal maintenance of VEGFR2 levels is required for transcriptional responses that determine cell fate. The time-frame in which most of the VEGFR2 degradation occurs in the absence of Rab5C, is also the time-frame that is necessary for VEGF signaling and the induction of immediate VEGF-responsive genes. Because Rab5C depletion leads to significant and quantitative differences in the expression of most VEGF-induced genes, our observations imply that accelerated VEGFR2 degradation results in disrupted VEGF-induced gene expression. Thus, the magnitude of VEGF-induced gene expression is closely related to VEGFR2 levels, and stabilization of endosomal VEGFR2 levels by Rab5C ensures the build-up of a sufficiently robust VEGF signal needed to surpass the threshold for transcription. Similar mechanisms likely regulate the duration of signaling of other receptors in different contexts, which is particularly interesting given the frequent overexpression of several Rab GTPases in cancer and many other diseases [54].
Third, the reduction in VEGFR2 levels induced by depletion of Rab5C is functionally and physiologically crucial, since we observe functional defects in tip cell behavior and angiogenic sprouting, both in vitro and in vivo. Tip cells are characterized by high VEGFR2 levels and thus high VEGF signaling, leading to the expression of tip cell markers [57]. Because VEGFR2 is downregulated in Rab5C-depleted cells whereas VEGFR1 levels are not affected, the VEGFR1/VEGFR2 balance is disturbed which confers a competitive disadvantage to develop tip cell properties [1113]. Indeed, the expression of various tip cell genes was decreased by Rab5C depletion, and tip cell formation was impaired in mosaic sprouting assays in vitro. While it is known that balanced VEGFR2 signaling regulates tip cells, we show that this crucially depends on endosomal regulation of VEGFR2 in a Rab5C-dependent manner, and that Rab5C deficiency is sufficient to affect all aspects of VEGF-driven tip cell responses, namely tip cell specification, function, and characteristics. Interestingly, out of eight different Rab5 GEFs, only the depletion of RIN2 specifically recapitulates our main findings in Rab5C-depleted cells. Although RIN2 may activate several Rabs and Rab5C might use multiple GEFs for all of its functions, our data indicate that RIN2 is the most important Rab5 GEF for the effects of RabC on VEGFR2 traffic and sprouting angiogenesis. These results nicely complement previous work on RIN2 in endothelial cells, where RIN2 was found to be important for several processes in endothelial cells including tube formation [55]. In line with previous reports of a high molecular and functional conservation of zebrafish and mammalian angiogenesis [46, 47], we could show full conservation of the Rin2/Rab5c-dependent Vegfr2-signaling axis in zebrafish, allowing us to analyze tip cell behavior and sprouting angiogenesis in response to Rin2/Rab5c-driven endosomal Vegfr2 signaling in vivo. In zebrafish embryos, depletion of Rab5c or Rin2 attenuated Vegf-induced activation of Notch signaling in the tip cells, filopodia formation, and endothelial cell migration, thus affecting key aspects of sprouting angiogenesis. Together, our data suggest that RIN2/Rab5C control a feedforward mechanism whereby inhibition of VEGFR2 degradation permits sprouting angiogenesis, by sustaining VEGF signaling and the expression of VEGF target genes. The endolysosomal machinery is therefore an important driver of the dynamic regulation of tip and stalk cells, and likely contributes strongly to their rapid and frequent interconversion observed during sprouting [56].
To date, few factors are known that protect VEGFR2 from degradation. Depletion of Numb or Rab GTPase-binding effector protein-2 (RABEP-2) also enhances VEGFR2 degradation, resulting in lower VEGFR2 levels and reduced VEGF signaling [27, 33]. Interestingly, genetic deletion of RABEP-2 in mice disturbs arteriogenesis but not angiogenesis, while the deletion of Numb induces angiogenic defects [27, 57]. It remains to be determined how different proteins along trafficking pathways can mediate such very diverse outcomes of VEGFR2 signaling. Future work should further establish how the fate of internalized VEGFR2 is linked to the transcriptional output of VEGF signaling and the dynamic regulation of endothelial cell phenotypes, as well as map the underlying molecular machinery along trafficking routes.

Materials and methods

Antibodies, plasmids and other materials

The following antibodies were used: anti-VEGFR2 (R&D Systems, AF357), anti-α-tubulin (Sigma-Aldrich, T6199), anti-AKT (Cell signaling technology, #9272), anti-p(S473)AKT (Cell signaling technology, #4060S), anti-VEGFR1 (Abcam, ab32152), anti-p(Y204)ERK-1/2 (Santa Cruz Biotechnology, sc-7383), anti-ERK-1/2 (Santa Cruz Biotechnology, sc-153), and anti-RIN2 (Sigma-Aldrich, HPA034641). HRP-conjugated secondary antibodies (#P0447, #P0399 and #P0449) were from DAKO, Alexa Fluor-conjugated secondary antibodies (A21445, A21200, and A21467) were from Thermo Fisher. Recombinant human VEGF-A (VEGF165) was purchased from R&D Systems (293-VE), fibronectin (#F1141) and thrombin (#T1063) were from Sigma-Aldrich, human fibrinogen (Haemocomplettan P, B02BB01) was from CSL Behring. Bafilomycin (#tlrl-baf1) was from InvivoGen, Leupeptin (#4041) and Pepstatin (#4397) were from PeptaNova, Hoechst 33,342 (#H3570), phalloidin (#T7471), and CellTracker Fluorescent Probes (CMFDA green, #C2925 and CMTPX red, #C34552) were from Thermo Fisher Scientific. All used primers were purchased from Integrated DNA Technologies, Inc. (Table S3). The human WT and DN RAB5C open reading frames (ORFs) were synthesized, cloned into pDONR221, and sequence-verified by Thermo Fisher Scientific for gateway cloning. For the zebrafish experiments, the RAB5C ORFs were cloned into a modified version of pTol2-fli1ep:CherryDest (a kind gift from Nathan Lawson; Addgene plasmid #73,493). To generate this destination vector, the mCherry sequence was PCR-amplified without stop codon using oligos extended with restriction sites (fwd: 5′-GACGATGCTAGCGCCACCATGGTGAGCAAGGGCGAG-3′ rev: 5′-GACGATTACGTAGATATCCTTGTACAGCTCGTCCATG-3′). This product was cloned into pTol2-fli1ep:CherryDest using the NheI and SnaBI restriction sites. Next, a Gateway Reading Frame Cassette A containing the appropriate attR1/2 sites was cloned into the EcoRV site, 3′ of the mCherry ORF, creating pTol2-fli1ep:CherryDestV2. Finally, the human WT and DN RAB5C ORFs in pDONR221 were cloned into pTol2-fli1ep:CherryDestV2 using LR Clonase II enzyme mix (Invitrogen), creating the 4 pTol2-fli1ep:Cherry-RAB5 plasmids. For lentiviral transduction of the RAB5 ORFs into HUVECs, we modified pLenti6.3/V5-Dest (Invitrogen) by removing an attR1/2 cassette by digestion with EcoRI and EcoRV, and introducing an NheI site using 2 oligos (5′-AATTCGCTAGCGACGACGAT-3′ and 5′-ATCGTCGTCGCTAGCG-3′). Next, the NheI/SnaBI fragment from pTol2-fli1ep:Cherry-RAB was cloned into the NheI/EcoRV sites of pLenti6.3/V5-Dest. To generate the mScarlet-Rab5C-Q80L construct, site-directed mutagenesis was performed on an mScarlet-Rab5C-wt construct using primer 5′-GTTTGAGATCTGGGACACAGCTGGACTGGAGCGGTATCACAGC-3′ and its reverse complement, in which CAG, coding for Q (Glutamine) was replaced with CTG coding for L (Leucine). All cloned plasmids used in experiments were sequenced with at least double coverage.

Cell culture, lentiviral transduction, and RNA interference

Primary HUVECs pooled from 3 to 5 individual donors (Lonza, C2519A) were cultured in endothelial growth medium-2 (EGM-2; Promocell, C-22011) supplemented with 2 mM L-glutamine (Sigma-Aldrich) and 100 U/ml penicillin and 100 μg/ml streptomycin (Sigma-Aldrich). Cell culture flasks and dishes were coated with 0.1% (w/v) gelatin (Sigma-Aldrich) unless stated otherwise. HUVECs were used between passages 3 and 6. To target Rab5C or the indicated GEFs, we used shRNAs cloned into pLKO.1 (3–9 per gene, Table S4) from the TRC Mission Library (a generous gift from Roderick Beijersbergen, The Netherlands Cancer Institute, Amsterdam). Human embryonic kidney (HEK) 293 T cells (ATCC, CRL-3216) were maintained in Dulbecco’s modified Eagle medium (DMEM) (Thermo Fisher Scientific) containing 4.5 g/l D-glucose, 2 mM L-glutamine, 10% (v/v) fetal bovine serum (Bodinco), 1 mM sodium pyruvate (Thermo Fisher Scientific), and 100 U/ml penicillin and 100 μg/ml streptomycin. To produce lentiviral particles containing shRNAs, HEK293T cells were transfected using TransIT-LT1 transfection reagent (Mirus Bio) according to the manufacturer’s protocol. Supernatant was harvested 48 and 72 h after transfection, centrifuged, filtered over a 0.45 μm pore filter, aliquoted and stored at -80ºC. HUVECs were lentivirally transduced with either a pool of shRNAs, or with a scrambled sequence in pLKO.1 as a negative control. Positive cells were selected during 3 days using 1 μg/ml puromycin (Sigma-Aldrich).

qPCR

For analysis of relative mRNA expression levels in transduced HUVECs, total RNA was isolated using the RNeasy kit (Qiagen) according to the manufacturer’s instructions. After reverse transcription to cDNA using the SuperScript III First-Strand Synthesis System (Thermo Fisher Scientific), qPCR was performed with the SensiFAST SYBR No-ROX kit (Bioline) and the indicated primers (Table S3), either on a LightCycler PCR system (Roche) or a StepOnePlus system (Applied Biosystems). Duplicate reactions were performed for each gene and expression was normalized to that of β-actin.

Flow cytometry

HUVECs were treated as indicated, detached with accutase (Sigma-Aldrich) for 2 min, and washed in 2% FCS in PBS. Cells were stained with primary antibodies for 1 h on ice, washed twice, followed by incubation with secondary antibodies for 45 min on ice and washed twice. Cells were analyzed on a Canto-II flow cytometer (BD Immunocytometry Systems) equipped with FACSDiva software.

Bio-informatic analysis of mRNA expression

Expression of Rab5 GEFs in HUVECs (Table S2) was determined by analysis of Affymetrix U133 Plus 2.0 genome-wide mRNA expression profiles in the public domain using the NCBI Gene Expression Omnibus (GEO: GSE7307) website [58, 59]. GEO was first searched for studies on low-passage, non-recombinant, non-stimulated HUVECs with data normalization using the MAS5.0 algorithm (Affymetrix Inc., Santa Barbara, CA). This resulted in a total of 11 published studies comprising 29 separate arrays, 17 of which were listed with present call analysis, that were queried for the expression of all known human Rab5 GEFs [20]. RINL was not represented on the Affymetrix arrays and was not further analyzed. Array data were analyzed as described using R2; an in-house developed Affymetrix analysis and visualization platform (http://​r2.​amc.​nl). Probes were ranked depending on high expression values and widespread expression (% of samples with significant expression for that gene) in the data collections tested.

Western blotting

Cells were washed in ice-cold PBS and lysed on ice in NP40 lysis buffer (50 mM Tris–HCl, pH 7.4, 100 mM NaCl, 10 mM MgCl2, 1% NP-40, 10% Glycerol), supplemented with protease inhibitor cocktail (Sigma-Aldrich). Cell lysates were cleared by centrifugation, heated at 95ºC in SDS sample buffer (50 mM Tris–HCl pH 6.8, 2% SDS, 10% glycerol, 1% β-mercaptoethanol, 12.5 mM EDTA, 0.02% bromophenol blue), and proteins were resolved by SDS-PAGE. Thereafter, proteins were transferred to nitrocellulose membranes (GE Healthcare Amersham) and aspecific binding was blocked using 5% (w/v) milk (Campina) in TBST (150 mM NaCl, 10 mM Tris, 0.1% Tween 20, pH 8.0) for 30 min. Immunoblotting was performed with the indicated antibodies by incubation with primary antibodies o/n at 4ºC and with secondary antibodies for 1 h at RT. Membranes were washed 3 × with TBST after each step. Proteins were visualized using ECL chemiluminescence reagent (Thermo Fisher Scientific), light sensitive films (Fuji Film) and a film processor (Konica Minolta, SRX-101A). Quantification of bands was performed by densitometry using ImageJ. Band intensity of the protein of interest was corrected to that of α-tubulin.

Degradation and signaling experiments

For VEGF signaling and VEGFR2 degradation assays, HUVECs in 6-well plates were grown to confluence. Then, cells were starved overnight and subsequently either left untreated or stimulated with 50 ng/ml VEGF for the indicated time-points, whereafter the cells were processed for Western blotting. Inhibitors were used as indicated.

Sprouting assays

Sprouting assays were performed according to previously established protocols with minor modifications [42, 43]. In brief, HUVECs were incubated with collagen-coated microcarrier beads (Sigma-Aldrich, C3275). The next day, the beads were detached by washing and transferred to 48-well plates containing fibrinogen in PBS (2 mg/ml) supplemented with 0.15 U/ml aprotinin and 6.25 U/ml thrombin. EGM-2 medium was added on top of the gels. Beads were imaged by time-lapse microscopy on a Widefield system using a 10 × dry lens objective (Carl Zeiss MicroImaging, Inc.). Images were recorded at 15-min intervals during 48 h. For each condition, 20 beads were analyzed per experiment. Sprouting was assessed as the average number of large (length of sprout > than the bead diameter) sprouts per bead. Alternatively, beads were embedded in fibrin gels in ‘half-area’ glass-bottom 96-well imaging plates (Corning, #4580). After 48 h, beads were fixed, stained with Hoechst and phalloidin, and visualized by generating z-stacks on a confocal microscope. Maximum projections of z-stacks were created using Fiji/ImageJ (version 1.52e), de-speckled once, and the contrast was enhanced with 0.3% and analyzed using the Angiogenesis plugin [60]. The phalloidin staining was used for correct segmentation of the sprouts. For the sprouts stained with CellTracker, the number of tip cells per condition was counted and normalized to the total number of cells in that condition.

Zebrafish husbandry and strains

Zebrafish (Danio rerio) husbandry and embryo maintenance were carried out under standard conditions at 28.5 ℃ [61]. Embryonic developmental stages were determined as described [62]. Transgenic lines used in this study were Tg(kdrl:GFP)s843, Tg(kdrl:Hras-mCherry)s896, Tg(TP1bglob:VenusPEST)s940, Tg(fli1a:lifeact-EGFP)mu240 [48, 51, 63, 64]. All animal experiments were performed in compliance with the relevant laws and institutional guidelines, were approved by local animal ethics committees and were conducted at the University of Münster and the MPI for Molecular Biomedicine with permissions granted by the Landesamt für Natur, Umwelt und Verbraucherschutz (LANUV) of North Rhine-Westphalia.

Generation of transgenic and rab5c mutant lines

The transgenic lines Tg(fli1a:mCherry-WT-hRAB5C)mu227 and Tg(fli1a:mCherry-DN-hRAB5C)mu228 zebrafish lines were newly generated using Tol2-mediated DNA integration [65]. The rab5cmu229 mutation was newly generated by CRISPR/Cas9. Annealed template oligonucleotides were transcribed into gRNAs using MEGAshortscript T7 Kit (Ambion) as described [66]. The following oligos were used for gRNA generation: rab5c 5′-AAAGCACCGACTCGGTGCCACTTTTTCAAGTTGATAACGGACTAGCCTTATTTTAACTTGCTATTTCTAGCTCTAAAACCCGTCGGGAACAAAATCTGCCTATAGTGAGTCGTATTACGC-3′ (rab5c target sequence: 5′-GCAGATTTTGTTCCCGACGGGGG). A total of 1 nl of 300 ng/ml gRNA and 3.22 μg/μl Cas9 NLS protein (M0646T, New England Biolabs) were injected at the single-cell stage. Efficiency of the CRISPR-targeting as well as genotyping were performed by PCR, using fwd 5′-GCCTCTCCATCCTTTTCTCA-3′ and rev 5′-TCACGAACACACTCCAGCTC-3′ and subsequent restriction enzyme digest with Hpy99I.

MOs and microinjections

For MO-mediated gene knockdown, embryos were injected at the one-cell stage with 3 ng control MO (Gene Tools), 5′-CCTCTTACCTCAGTTACAATTTATA-3′, 3 ng rab5c MO, 5′-CGCTGGTCCACCTCGCCCCGCCATG-3′ 3 ng rin2 MO, 5′-GCAGTGTGTTTTAACTCTCACCTTA-3′. Sequence of the rab5c ATG MO was as previously described [37]. The rab5c and rin2 splice-site MOs were generated as shown (Fig. S6A,S8A), and validated by RT-PCR. For mosaic overexpression of mCherry-WT-Rab5C or mCherry-DN-Rab5C, embryos were injected at the single-cell stage with 100 pg plasmid and 250 pg Tol2 mRNA per embryo. Effects on vascular development were analyzed at 30–32 hpf.

RT-PCR and Western blotting in zebrafish

Total RNAs were prepared from 36 hpf embryos using TRIzol reagent (Invitrogen), and reverse-transcribed by random hexamer primers using Superscript III (Invitrogen) according to the manufacturer’s instruction. PCR was performed using gene-specific primers as below. Efficiency of the knockdown was assessed by PCR using the primers fwd: AGAGCTGCGTACATCCAACC and rev: TGCAAGCCAAAATTCAAAGA.
After removing chorionic membrane and yolk sac, embryos injected with either control MO or rab5c MO were directly lysed in 1 × SDS sample buffer, and subjected to Western blot analysis with anti-Kdrl antibody (Kerafast, ES1003) and anti-alpha-Actin antibody (Santa Cruz Biotechnology, sc-47778) as described previously [67].

Confocal microscopy

HUVECs on glass coverslips were fixed with 4% paraformaldehyde (Merck) in PBS containing 1 mM CaCl2 and 0.5 mM MgCl2 (PBS + +) for 10 min, and permeabilized with 0.4% Triton X-100 (Sigma-Aldrich) in PBS +  + for 5 min. Aspecific antibody binding was prevented by blocking with 2% BSA Fraction V (Sigma-Aldrich) in PBS +  + for 15 min. Following incubation with the indicated primary antibodies, coverslips were washed with 0.5% BSA Fraction V in PBS +  + and antibody binding was visualized using secondary antibodies. Coverslips were washed and subsequently mounted in 10% Mowiol, 2.5% Dabco, 25% glycerol, pH 8.5. Image acquisition was performed on a Leica SP8 confocal microscope using a 63 × oil immersion objective, after which images were processed using Leica Application Suite X and Fiji/ImageJ (version 1.52e) software.
Sprouting was visualized using a 25 × long-distance water objective on a Leica SP5 confocal set-up. z-stacks were created of 1.5 μm per slice and the different channels were sequentially scanned between stacks.
Confocal microscopy of zebrafish was performed using an LSM780 (Carl Zeiss Microscopy GmbH; objective lens: Plan Achromat 20/0.8, 20x; LD C-Apochromat 40x/1.1 W Korr M27). Living zebrafish embryos were embedded in 1% agarose (dissolved in E3 medium containing Tricaine) and analyzed using Imaris 9 software (Bitplane) as previously described [68].
For time-lapse analysis in zebrafish, a temperature of 28.5 °C was maintained using a heating chamber. Images were collected every 6 min during 1 h (for imaging of filopodia) or every 15 min during 6 h (for imaging of ISV sprouting). Assembly of confocal stacks and time-lapse movies was conducted using Imaris 8/9 software (Bitplane).

mRNA sequencing and analysis

sh_Ctrl, sh_Rab5A or sh_Rab5C cells were seeded in 10 cm dishes and grown to confluence. Cells were starved overnight, whereafter they were either left untreated or stimulated with 50 ng/ml VEGF for 1 h or 4 h. Subsequently, cells were lysed using RLT buffer (Qiagen) with 1% β-mercapto-ethanol and stored at −80 °C until analysis. RNA-seq libraries were prepared from samples of 2 independent experiments with the mRNA KAPA HyperPrep Kit (Illumina) using Truseq Illumina LT adapters. Sequencing was performed on an Illumina Hiseq4000 system (single-read 50 bp), 10 samples per lane.
Reads were subjected to quality control (FastQC, dupRadar, Picard Tools), trimmed using Trimmomatic v0.36, and aligned to the hg38 genome using HISAT2 v2.1.0. Counts were obtained with HTSeq v0.11.0. Statistical analyses were performed using the edgeR and limma/voom packages [69, 70]. Genes with no counts in any of the samples were removed while genes with more than 2 reads in at least 3 of the samples were kept. Count data were transformed to log2-counts per million (logCPM), normalized by applying the trimmed mean of M-values method and precision-weighted using voom. Differential expression was assessed using an empirical Bayes’ moderated t-test within limma’s linear model framework including the precision weights. Resulting p values were corrected for multiple testing using the Benjamini–Hochberg false discovery rate. Additional gene annotation was retrieved from Ensembl (release 94) using the biomaRt/Bioconductor package. Geneset enrichment was performed using CAMERA with a preset value of 0.01 for the inter-gene correlation using collections H, C1, C2, C3, C5, C6, and C7 retrieved from the Molecular Signatures Database (MSigDB v6.2; Entrez Gene ID version), complemented with a user-defined geneset containing the genes constituting the observed VEGF transcriptome. Data analysis were performed using R (v3.5.2) and Bioconductor (v3.8). Sequencing data have been made available in the GEO under accession number: GSE134947.

Statistical analysis

Statistical analysis was performed using one-way ANOVA for multiple comparisons and unpaired t-tests for comparisons between two conditions, unless stated otherwise. Throughout the paper, statistically significant differences are indicated by * (p < 0.05), ** (p < 0.01), *** (p < 0.001), or **** (p < 0.0001).

Acknowledgements

We thank Arnoud Sonnenberg (Department of Cell Biology, Netherlands Cancer Institute, Amsterdam), Ben Morris and Roderick Beijersbergen (Robotics and Screening Center, Netherlands Cancer Institute, Amsterdam) and Nathan Lawson (UMass Medical School, Worcester, MA) for their kind gifts of materials. We gratefully acknowledge technical support from Paul Verkuijlen (Sanquin Research, Amsterdam), Simon Tol, Mark Hoogenboezen, and Erik Mul (Central Facility, Sanquin Research, Amsterdam), Lenny Brocks and Marjolijn Mertz (Digital microscopy facility, Netherlands Cancer Institute, Amsterdam), and Linda Koster (Sequencing core facility, Amsterdam University Medical Centers, Amsterdam). We thank Arnoud Sonnenberg for critical reading. This work was supported by research grants from the Netherlands Organisation for Scientific Research (ZonMW Veni 016.146.160) and the Dutch Thrombosis Foundation (2017-01) awarded to CM, Sanquin (PPOC 14-43) awarded to DdK, the Japan Society for the Promotion of Science Overseas Researcher Fellowships awarded to YW, and the Deutsche Forschungsgemeinschaft (HE4585/3-1), awarded to WH. We further acknowledge support from the Dutch Society for Cell Biology.

Declarations

Conflict of interest

The authors declare that they have no conflict of interest.
Open AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher's Note

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

Unsere Produktempfehlungen

Die Chirurgie

Print-Titel

Das Abo mit mehr Tiefe

Mit der Zeitschrift Die Chirurgie erhalten Sie zusätzlich Online-Zugriff auf weitere 43 chirurgische Fachzeitschriften, CME-Fortbildungen, Webinare, Vorbereitungskursen zur Facharztprüfung und die digitale Enzyklopädie e.Medpedia.

Bis 30. April 2024 bestellen und im ersten Jahr nur 199 € zahlen!

e.Med Interdisziplinär

Kombi-Abonnement

Für Ihren Erfolg in Klinik und Praxis - Die beste Hilfe in Ihrem Arbeitsalltag

Mit e.Med Interdisziplinär erhalten Sie Zugang zu allen CME-Fortbildungen und Fachzeitschriften auf SpringerMedizin.de.

e.Med Innere Medizin

Kombi-Abonnement

Mit e.Med Innere Medizin erhalten Sie Zugang zu CME-Fortbildungen des Fachgebietes Innere Medizin, den Premium-Inhalten der internistischen Fachzeitschriften, inklusive einer gedruckten internistischen Zeitschrift Ihrer Wahl.

Literatur
2.
Zurück zum Zitat Simons M, Gordon E, Claesson-Welsh L (2016) Mechanisms and regulation of endothelial VEGF receptor signalling. Nat Rev Mol Cell Biol 17:611–625PubMedCrossRef Simons M, Gordon E, Claesson-Welsh L (2016) Mechanisms and regulation of endothelial VEGF receptor signalling. Nat Rev Mol Cell Biol 17:611–625PubMedCrossRef
4.
Zurück zum Zitat Gerhardt H, Golding M, Fruttiger M, Ruhrberg C, Lundkvist A, Abramsson A, Jeltsch M, Mitchell C, Alitalo K, Shima D, Betsholtz C (2003) VEGF guides angiogenic sprouting utilizing endothelial tip cell filopodia. J Cell Biol 161:1163–1177PubMedPubMedCentralCrossRef Gerhardt H, Golding M, Fruttiger M, Ruhrberg C, Lundkvist A, Abramsson A, Jeltsch M, Mitchell C, Alitalo K, Shima D, Betsholtz C (2003) VEGF guides angiogenic sprouting utilizing endothelial tip cell filopodia. J Cell Biol 161:1163–1177PubMedPubMedCentralCrossRef
5.
Zurück zum Zitat Herbert S, Stainier D (2011) Molecular control of endothelial cell behaviour during blood vessel morphogenesis. Nat Rev Mol Cell Biol 12:551–564PubMedPubMedCentralCrossRef Herbert S, Stainier D (2011) Molecular control of endothelial cell behaviour during blood vessel morphogenesis. Nat Rev Mol Cell Biol 12:551–564PubMedPubMedCentralCrossRef
6.
Zurück zum Zitat Strasser G, Kaminker J, Tessier-Lavigne M (2010) Microarray analysis of retinal endothelial tip cells identifies CXCR4 as a mediator of tip cell morphology and branching. Blood 115:5102–5110PubMedCrossRef Strasser G, Kaminker J, Tessier-Lavigne M (2010) Microarray analysis of retinal endothelial tip cells identifies CXCR4 as a mediator of tip cell morphology and branching. Blood 115:5102–5110PubMedCrossRef
7.
Zurück zum Zitat del Toro R, Prahst C, Mathivet T, Siegfried G, Kaminker J, Larrivee B, Breant C, Duarte A, Takakura N, Fukamizu A, Penninger J, Eichmann A (2010) Identification and functional analysis of endothelial tip cell-enriched genes. Blood 116:4025–4033PubMedPubMedCentralCrossRef del Toro R, Prahst C, Mathivet T, Siegfried G, Kaminker J, Larrivee B, Breant C, Duarte A, Takakura N, Fukamizu A, Penninger J, Eichmann A (2010) Identification and functional analysis of endothelial tip cell-enriched genes. Blood 116:4025–4033PubMedPubMedCentralCrossRef
8.
Zurück zum Zitat Hellstrom M, Phng L, Hofmann J, Wallgard E, Coultas L, Lindblom P, Alva J, Nilsson A-K, Karlsson L, Gaiano N, Yoon K, Rossant J, Iruela-Arispe L, Kalén M, Gerhardt H, Betsholtz C (2007) Dll4 signalling through Notch1 regulates formation of tip cells during angiogenesis. Nature 445:776–780CrossRefPubMed Hellstrom M, Phng L, Hofmann J, Wallgard E, Coultas L, Lindblom P, Alva J, Nilsson A-K, Karlsson L, Gaiano N, Yoon K, Rossant J, Iruela-Arispe L, Kalén M, Gerhardt H, Betsholtz C (2007) Dll4 signalling through Notch1 regulates formation of tip cells during angiogenesis. Nature 445:776–780CrossRefPubMed
9.
Zurück zum Zitat Lobov I, Renard R, Papadopoulos N, Gale N, Thurston G, Yancopoulos G, Wiegand S (2007) Delta-like ligand 4 (Dll4) is induced by VEGF as a negative regulator of angiogenic sprouting. Proc Natl Acad Sci USA 104:3219–3224PubMedPubMedCentralCrossRef Lobov I, Renard R, Papadopoulos N, Gale N, Thurston G, Yancopoulos G, Wiegand S (2007) Delta-like ligand 4 (Dll4) is induced by VEGF as a negative regulator of angiogenic sprouting. Proc Natl Acad Sci USA 104:3219–3224PubMedPubMedCentralCrossRef
10.
Zurück zum Zitat Suchting S, Freitas C, le Noble F, Benedito R, Bréant C, Duarte A, Eichmann A (2007) The Notch ligand Delta-like 4 negatively regulates endothelial tip cell formation and vessel branching. Proc Natl Acad Sci U S A 104:3225–3230PubMedPubMedCentralCrossRef Suchting S, Freitas C, le Noble F, Benedito R, Bréant C, Duarte A, Eichmann A (2007) The Notch ligand Delta-like 4 negatively regulates endothelial tip cell formation and vessel branching. Proc Natl Acad Sci U S A 104:3225–3230PubMedPubMedCentralCrossRef
11.
Zurück zum Zitat Fong G, Rossant J, Gertsenstein M, Breitman M (1995) Role of the Flt-1 receptor tyrosine kinase in regulating the assembly of vascular endothelium. Nature 376:66–70PubMedCrossRef Fong G, Rossant J, Gertsenstein M, Breitman M (1995) Role of the Flt-1 receptor tyrosine kinase in regulating the assembly of vascular endothelium. Nature 376:66–70PubMedCrossRef
12.
Zurück zum Zitat Hiratsuka S, Minowa O, Kuno J, Noda T, Shibuya M (1998) Flt-1 lacking the tyrosine kinase domain is sufficient for normal development and angiogenesis in mice. Proc Natl Acad Sci USA 95:9349–9354PubMedPubMedCentralCrossRef Hiratsuka S, Minowa O, Kuno J, Noda T, Shibuya M (1998) Flt-1 lacking the tyrosine kinase domain is sufficient for normal development and angiogenesis in mice. Proc Natl Acad Sci USA 95:9349–9354PubMedPubMedCentralCrossRef
13.
Zurück zum Zitat Krueger J, Liu D, Scholz K, Zimmer A, Shi Y, Klein C, Siekmann A, Schulte-Merker S, Cudmore M, Ahmed A, le Noble F (2011) Flt1 acts as a negative regulator of tip cell formation and branching morphogenesis in the zebrafish embryo. Development 138:2111–2120PubMedPubMedCentralCrossRef Krueger J, Liu D, Scholz K, Zimmer A, Shi Y, Klein C, Siekmann A, Schulte-Merker S, Cudmore M, Ahmed A, le Noble F (2011) Flt1 acts as a negative regulator of tip cell formation and branching morphogenesis in the zebrafish embryo. Development 138:2111–2120PubMedPubMedCentralCrossRef
14.
Zurück zum Zitat Ballmer-Hofer K, Andersson A, Ratcliffe L, Berger P (2011) Neuropilin-1 promotes VEGFR-2 trafficking through Rab11 vesicles thereby specifying signal output. Blood 118:816–826PubMedCrossRef Ballmer-Hofer K, Andersson A, Ratcliffe L, Berger P (2011) Neuropilin-1 promotes VEGFR-2 trafficking through Rab11 vesicles thereby specifying signal output. Blood 118:816–826PubMedCrossRef
15.
Zurück zum Zitat Basagiannis D, Zografou S, Murphy C, Fotsis T, Morbidelli L, Ziche M, Bleck C, Mercer J, Christoforidis S (2016) VEGF induces signalling and angiogenesis by directing VEGFR2 internalisation through macropinocytosis. J Cell Sci 129:4091–4104PubMed Basagiannis D, Zografou S, Murphy C, Fotsis T, Morbidelli L, Ziche M, Bleck C, Mercer J, Christoforidis S (2016) VEGF induces signalling and angiogenesis by directing VEGFR2 internalisation through macropinocytosis. J Cell Sci 129:4091–4104PubMed
16.
Zurück zum Zitat Gampel A, Moss L, Jones M, Brunton V, Norman J, Mellor H (2006) VEGF regulates the mobilization of VEGFR2/KDR from an intracellular endothelial storage compartment. Blood 108:2624–2631PubMedCrossRef Gampel A, Moss L, Jones M, Brunton V, Norman J, Mellor H (2006) VEGF regulates the mobilization of VEGFR2/KDR from an intracellular endothelial storage compartment. Blood 108:2624–2631PubMedCrossRef
17.
Zurück zum Zitat Jopling H, Howell G, Gamper N, Ponnambalam S (2011) The VEGFR2 receptor tyrosine kinase undergoes constitutive endosome-to-plasma membrane recycling. Biochem Biophys Res Commun 410:170–176PubMedPubMedCentralCrossRef Jopling H, Howell G, Gamper N, Ponnambalam S (2011) The VEGFR2 receptor tyrosine kinase undergoes constitutive endosome-to-plasma membrane recycling. Biochem Biophys Res Commun 410:170–176PubMedPubMedCentralCrossRef
18.
Zurück zum Zitat Lampugnani M, Orsenigo F, Gagliani M, Tacchetti C, Dejana E (2006) Vascular endothelial cadherin controls VEGFR2 internalization and signaling from intracellular compartments. J Cell Biol 174:593–604PubMedPubMedCentralCrossRef Lampugnani M, Orsenigo F, Gagliani M, Tacchetti C, Dejana E (2006) Vascular endothelial cadherin controls VEGFR2 internalization and signaling from intracellular compartments. J Cell Biol 174:593–604PubMedPubMedCentralCrossRef
19.
Zurück zum Zitat Langemeyer L, Fröhlich F, Ungermann C (2018) Rab GTPase function in endosome and lysosome biogenesis. Trends Cell Biol 28:957–970PubMedCrossRef Langemeyer L, Fröhlich F, Ungermann C (2018) Rab GTPase function in endosome and lysosome biogenesis. Trends Cell Biol 28:957–970PubMedCrossRef
20.
Zurück zum Zitat Ishida M, Oguchi ME, Fukuda M (2016) Multiple types of guanine nucleotide exchange factors (GEFs) for Rab small GTPases. Cell Struct Funct 41:61–79PubMedCrossRef Ishida M, Oguchi ME, Fukuda M (2016) Multiple types of guanine nucleotide exchange factors (GEFs) for Rab small GTPases. Cell Struct Funct 41:61–79PubMedCrossRef
21.
22.
Zurück zum Zitat Ewan L, Jopling H, Jia H, Mittar S, Bagherzadeh A, Howell G, Walker J, Zachary I, Ponnambalam S (2006) Intrinsic tyrosine kinase activity is required for vascular endothelial growth factor receptor 2 ubiquitination, sorting and degradation in endothelial cells. Traffic 7:1270–1282PubMedCrossRef Ewan L, Jopling H, Jia H, Mittar S, Bagherzadeh A, Howell G, Walker J, Zachary I, Ponnambalam S (2006) Intrinsic tyrosine kinase activity is required for vascular endothelial growth factor receptor 2 ubiquitination, sorting and degradation in endothelial cells. Traffic 7:1270–1282PubMedCrossRef
23.
Zurück zum Zitat Nakayama M, Nakayama A, van Lessen M, Yamamoto H, Hoffmann S, Drexler H, Itoh N, Hirose T, Breier G, Vestweber D, Cooper J, Ohno S, Kaibuchi K, Adams R (2013) Spatial regulation of VEGF receptor endocytosis in angiogenesis. Nat Cell Biol 15:249–260PubMedPubMedCentralCrossRef Nakayama M, Nakayama A, van Lessen M, Yamamoto H, Hoffmann S, Drexler H, Itoh N, Hirose T, Breier G, Vestweber D, Cooper J, Ohno S, Kaibuchi K, Adams R (2013) Spatial regulation of VEGF receptor endocytosis in angiogenesis. Nat Cell Biol 15:249–260PubMedPubMedCentralCrossRef
24.
Zurück zum Zitat Rahman H, Wu H, Dong Y, Al E (2016) Selective targeting of a novel epsin-VEGFR2 interaction promotes VEGF-mediated angiogenesis. Circ Res 118:957–969PubMedPubMedCentralCrossRef Rahman H, Wu H, Dong Y, Al E (2016) Selective targeting of a novel epsin-VEGFR2 interaction promotes VEGF-mediated angiogenesis. Circ Res 118:957–969PubMedPubMedCentralCrossRef
25.
Zurück zum Zitat Tessneer K, Pasula S, Cai X, Dong Y, McManus J, Liu X, Yu L, Hahn S, Chang B, Chen Y, Griffin C, Xia L, Adams R, Chen H (2014) Genetic reduction of vascular endothelial growth factor receptor 2 rescues aberrant angiogenesis caused by epsin deficiency. Arter Thromb Vasc Biol 34:331–337CrossRef Tessneer K, Pasula S, Cai X, Dong Y, McManus J, Liu X, Yu L, Hahn S, Chang B, Chen Y, Griffin C, Xia L, Adams R, Chen H (2014) Genetic reduction of vascular endothelial growth factor receptor 2 rescues aberrant angiogenesis caused by epsin deficiency. Arter Thromb Vasc Biol 34:331–337CrossRef
26.
Zurück zum Zitat Lanahan A, Zhang X, Fantin A, Zhuang Z, Rivera-Molina F, Speichinger K, Prahst C, Zhang J, Wang Y, Davis G, Toomre D, Ruhrberg C, Simons M (2013) The neuropilin-1 cytoplasmic domain is required for VEGF-A-dependent arteriogenesis. Dev Cell 25:156–168PubMedPubMedCentralCrossRef Lanahan A, Zhang X, Fantin A, Zhuang Z, Rivera-Molina F, Speichinger K, Prahst C, Zhang J, Wang Y, Davis G, Toomre D, Ruhrberg C, Simons M (2013) The neuropilin-1 cytoplasmic domain is required for VEGF-A-dependent arteriogenesis. Dev Cell 25:156–168PubMedPubMedCentralCrossRef
27.
Zurück zum Zitat van Lessen M, Nakayama M, Kato K, Kim J, Kaibuchi K, Adams R (2015) Regulation of vascular endothelial growth factor receptor function in angiogenesis by numb and numb-like. Arter Thromb Vasc Biol 35:1815–1825CrossRef van Lessen M, Nakayama M, Kato K, Kim J, Kaibuchi K, Adams R (2015) Regulation of vascular endothelial growth factor receptor function in angiogenesis by numb and numb-like. Arter Thromb Vasc Biol 35:1815–1825CrossRef
28.
Zurück zum Zitat Pasula S, Cai X, Dong Y, Messa M, McManus J, Chang B, Liu X, Zhu H, Mansat RS, Yoon SJ, Hahn S, Keeling J, Saunders D, Ko G, Knight J, Newton G, Luscinskas F, Sun X, Towner R, Lupu F et al (2012) Endothelial epsin deficiency decreases tumor growth by enhancing VEGF signaling. J Clin Invest 122:4424–4438PubMedPubMedCentralCrossRef Pasula S, Cai X, Dong Y, Messa M, McManus J, Chang B, Liu X, Zhu H, Mansat RS, Yoon SJ, Hahn S, Keeling J, Saunders D, Ko G, Knight J, Newton G, Luscinskas F, Sun X, Towner R, Lupu F et al (2012) Endothelial epsin deficiency decreases tumor growth by enhancing VEGF signaling. J Clin Invest 122:4424–4438PubMedPubMedCentralCrossRef
29.
Zurück zum Zitat Reynolds A, Hart I, Watson A, Welti J, Silva R, Robinson S, Violante G, Gourlaouen M, Salih M, Jones M, Jones D, Saunders G, Kostourou V, Perron-Sierra F, Norman J, Tucker G, Hodivala-Dilke K (2009) Stimulation of tumor growth and angiogenesis by low concentrations of RGD-mimetic integrin inhibitors. Nat Med 15:392–400PubMedCrossRef Reynolds A, Hart I, Watson A, Welti J, Silva R, Robinson S, Violante G, Gourlaouen M, Salih M, Jones M, Jones D, Saunders G, Kostourou V, Perron-Sierra F, Norman J, Tucker G, Hodivala-Dilke K (2009) Stimulation of tumor growth and angiogenesis by low concentrations of RGD-mimetic integrin inhibitors. Nat Med 15:392–400PubMedCrossRef
30.
Zurück zum Zitat Sawamiphak S, Seidel S, Essmann C, Wilkinson G, Pitulescu M, Acker T, Acker-Palmer A (2010) Ephrin-B2 regulates VEGFR2 function in developmental and tumour angiogenesis. Nature 465:487–491PubMedCrossRef Sawamiphak S, Seidel S, Essmann C, Wilkinson G, Pitulescu M, Acker T, Acker-Palmer A (2010) Ephrin-B2 regulates VEGFR2 function in developmental and tumour angiogenesis. Nature 465:487–491PubMedCrossRef
31.
Zurück zum Zitat Wang Y, Nakayama M, Pitulescu M, Schmidt TS, Bochenek ML, Sakakibara A, Adams S, Davy A, Deutsch U, Lüthi U, Barberis A, Benjamin LE, Mäkinen T, Nobes CD, Adams RH (2010) Ephrin-B2 controls VEGF-induced angiogenesis and lymphangiogenesis. Nature 465:483–486PubMedCrossRef Wang Y, Nakayama M, Pitulescu M, Schmidt TS, Bochenek ML, Sakakibara A, Adams S, Davy A, Deutsch U, Lüthi U, Barberis A, Benjamin LE, Mäkinen T, Nobes CD, Adams RH (2010) Ephrin-B2 controls VEGF-induced angiogenesis and lymphangiogenesis. Nature 465:483–486PubMedCrossRef
32.
Zurück zum Zitat Lee MY, Skoura A, Park EJ, Landskroner-eiger S, Jozsef L, Luciano AK, Murata T, Pasula S, Dong Y, Bouaouina M, Calderwood DA, Ferguson SM, De Camilli P, Sessa WC (2014) Dynamin 2 regulation of integrin endocytosis, but not VEGF signaling, is crucial for developmental angiogenesis. Development 141(7):1465–1472PubMedPubMedCentralCrossRef Lee MY, Skoura A, Park EJ, Landskroner-eiger S, Jozsef L, Luciano AK, Murata T, Pasula S, Dong Y, Bouaouina M, Calderwood DA, Ferguson SM, De Camilli P, Sessa WC (2014) Dynamin 2 regulation of integrin endocytosis, but not VEGF signaling, is crucial for developmental angiogenesis. Development 141(7):1465–1472PubMedPubMedCentralCrossRef
33.
Zurück zum Zitat Kofler N, Corti F, Rivera-Molina F, Deng Y, Toomre D, Simons M (2018) The Rab-effector protein RABEP2 regulates endosomal trafficking to mediate vascular endothelial growth factor receptor-2 (VEGFR2)-dependent signaling. J Biol Chem 293:4805–4817PubMedPubMedCentralCrossRef Kofler N, Corti F, Rivera-Molina F, Deng Y, Toomre D, Simons M (2018) The Rab-effector protein RABEP2 regulates endosomal trafficking to mediate vascular endothelial growth factor receptor-2 (VEGFR2)-dependent signaling. J Biol Chem 293:4805–4817PubMedPubMedCentralCrossRef
34.
Zurück zum Zitat Lanahan A, Hermans K, Claes F, Kerley-Hamilton J, Zhuang Z, Giordano F, Carmeliet P, Simons M (2010) VEGF receptor 2 endocytic trafficking regulates arterial morphogenesis. Dev Cell 18:713–724PubMedPubMedCentralCrossRef Lanahan A, Hermans K, Claes F, Kerley-Hamilton J, Zhuang Z, Giordano F, Carmeliet P, Simons M (2010) VEGF receptor 2 endocytic trafficking regulates arterial morphogenesis. Dev Cell 18:713–724PubMedPubMedCentralCrossRef
35.
Zurück zum Zitat Clark B, Winter M, Cohen A, Link B (2011) Generation of Rab-based transgenic lines for in vivo studies of endosome biology in zebrafish. Dev Dyn 240:2452–2465PubMedPubMedCentralCrossRef Clark B, Winter M, Cohen A, Link B (2011) Generation of Rab-based transgenic lines for in vivo studies of endosome biology in zebrafish. Dev Dyn 240:2452–2465PubMedPubMedCentralCrossRef
36.
Zurück zum Zitat Heng J, Lv P, Zhang Y, Cheng X, Wang L, Ma D, Liu F (2020) Rab5c-mediated endocytic trafficking regulates hematopoietic stem and progenitor cell development via Notch and AKT signaling. PLoS Biol 18:1–28CrossRef Heng J, Lv P, Zhang Y, Cheng X, Wang L, Ma D, Liu F (2020) Rab5c-mediated endocytic trafficking regulates hematopoietic stem and progenitor cell development via Notch and AKT signaling. PLoS Biol 18:1–28CrossRef
37.
Zurück zum Zitat Kenyon E, Campos I, Bull J, Williams P, Stemple D, Clark M (2015) Zebrafish Rab5 proteins and a role for Rab5ab in nodal signalling. Dev Biol 397:212–224PubMedPubMedCentralCrossRef Kenyon E, Campos I, Bull J, Williams P, Stemple D, Clark M (2015) Zebrafish Rab5 proteins and a role for Rab5ab in nodal signalling. Dev Biol 397:212–224PubMedPubMedCentralCrossRef
38.
Zurück zum Zitat Thisse B, Thisse C (2004) Fast Release Clones: A High Throughput Expression Analysis. ZFIN Direct Data Submiss. http://zfin.org Thisse B, Thisse C (2004) Fast Release Clones: A High Throughput Expression Analysis. ZFIN Direct Data Submiss. http://​zfin.​org
39.
Zurück zum Zitat Fish J, Cantu Gutierrez M, Dang L, Khyzha N, Chen Z, Veitch S, Cheng H, Khor M, Antounians L, Njock M, Boudreau E, Herman A, Rhyner A, Ruiz O, Eisenhoffer G, Medina-Rivera A, Wilson M, Wythe J (2017) Dynamic regulation of VEGF-inducible genes by an ERK/ERG/p300 transcriptional network. Development 144:2428–2444PubMedPubMedCentral Fish J, Cantu Gutierrez M, Dang L, Khyzha N, Chen Z, Veitch S, Cheng H, Khor M, Antounians L, Njock M, Boudreau E, Herman A, Rhyner A, Ruiz O, Eisenhoffer G, Medina-Rivera A, Wilson M, Wythe J (2017) Dynamic regulation of VEGF-inducible genes by an ERK/ERG/p300 transcriptional network. Development 144:2428–2444PubMedPubMedCentral
40.
Zurück zum Zitat Schweighofer B, Testori J, Sturtzel C, Sattler S, Mayer H, Wagner O, Bilban M, Hofer E (2009) The VEGF-induced transcriptional response comprises gene clusters at the crossroad of angiogenesis and inflammation. Thromb Haemost 102:544–554PubMedPubMedCentralCrossRef Schweighofer B, Testori J, Sturtzel C, Sattler S, Mayer H, Wagner O, Bilban M, Hofer E (2009) The VEGF-induced transcriptional response comprises gene clusters at the crossroad of angiogenesis and inflammation. Thromb Haemost 102:544–554PubMedPubMedCentralCrossRef
41.
Zurück zum Zitat Shin J, Huggenberger R, Detmar M (2008) Transcriptional profiling of VEGF-A and VEGF-C target genes in lymphatic endothelium reveals endothelial-specific molecule-1 as a novel mediator of lymphangiogenesis. Blood 112:2318–2326PubMedPubMedCentralCrossRef Shin J, Huggenberger R, Detmar M (2008) Transcriptional profiling of VEGF-A and VEGF-C target genes in lymphatic endothelium reveals endothelial-specific molecule-1 as a novel mediator of lymphangiogenesis. Blood 112:2318–2326PubMedPubMedCentralCrossRef
42.
Zurück zum Zitat Nakatsu MN, Hughes CC (2008) An optimised three-dimensional in vitro model for the analysis of angiogenesis. Methods Enzym 443:65–82CrossRef Nakatsu MN, Hughes CC (2008) An optimised three-dimensional in vitro model for the analysis of angiogenesis. Methods Enzym 443:65–82CrossRef
43.
Zurück zum Zitat Nowak-Sliwinska P, Alitalo K, Allen E, Anisimov A, Aplin AC, Auerbach R, Augustin HG, Bates DO, van Beijnum JR, Bender RHF, Bergers G, Bikfalvi A, Bischoff J, Böck BC, Griffioen AW (2018) Consensus guidelines for the use and interpretation of angiogenesis assays. Angiogenesis 21:425–532PubMedPubMedCentralCrossRef Nowak-Sliwinska P, Alitalo K, Allen E, Anisimov A, Aplin AC, Auerbach R, Augustin HG, Bates DO, van Beijnum JR, Bender RHF, Bergers G, Bikfalvi A, Bischoff J, Böck BC, Griffioen AW (2018) Consensus guidelines for the use and interpretation of angiogenesis assays. Angiogenesis 21:425–532PubMedPubMedCentralCrossRef
45.
46.
Zurück zum Zitat Ellertsdóttir E, Lenard A, Blum Y, Krudewig A, Herwig L, Affolter M, Belting H (2010) Vascular morphogenesis in the zebrafish embryo. Dev Biol 341:56–65PubMedCrossRef Ellertsdóttir E, Lenard A, Blum Y, Krudewig A, Herwig L, Affolter M, Belting H (2010) Vascular morphogenesis in the zebrafish embryo. Dev Biol 341:56–65PubMedCrossRef
47.
Zurück zum Zitat Schuermann A, Helker C, Herzog W (2014) Angiogenesis in zebrafish. Semin Cell Dev Biol 31:106–114PubMedCrossRef Schuermann A, Helker C, Herzog W (2014) Angiogenesis in zebrafish. Semin Cell Dev Biol 31:106–114PubMedCrossRef
48.
Zurück zum Zitat Hamm M, Kirchmaier B, Herzog W (2016) Sema3d controls collective endothelial cell migration by distinct mechanisms via Nrp1 and PlxnD1. J Cell Biol 215:415–430PubMedPubMedCentralCrossRef Hamm M, Kirchmaier B, Herzog W (2016) Sema3d controls collective endothelial cell migration by distinct mechanisms via Nrp1 and PlxnD1. J Cell Biol 215:415–430PubMedPubMedCentralCrossRef
49.
Zurück zum Zitat Hasan S, Tsaryk R, Lange M, Wisniewski L, Moore J, Lawson N, Wojciechowska K, Schnittler H, Siekmann A (2017) Endothelial Notch signalling limits angiogenesis via control of artery formation. Nat Cell Biol 19:928–940PubMedPubMedCentralCrossRef Hasan S, Tsaryk R, Lange M, Wisniewski L, Moore J, Lawson N, Wojciechowska K, Schnittler H, Siekmann A (2017) Endothelial Notch signalling limits angiogenesis via control of artery formation. Nat Cell Biol 19:928–940PubMedPubMedCentralCrossRef
50.
Zurück zum Zitat Pitulescu M, Schmidt I, Giaimo B, Antoine T, Berkenfeld F, Ferrante F, Park H, Ehling M, Biljes D, Rocha S, Langen U, Stehling M, Nagasawa T, Ferrara N, Borggrefe T, Adams R (2017) Dll4 and Notch signalling couples sprouting angiogenesis and artery formation. Nat Cell Biol 19:915–927PubMedCrossRef Pitulescu M, Schmidt I, Giaimo B, Antoine T, Berkenfeld F, Ferrante F, Park H, Ehling M, Biljes D, Rocha S, Langen U, Stehling M, Nagasawa T, Ferrara N, Borggrefe T, Adams R (2017) Dll4 and Notch signalling couples sprouting angiogenesis and artery formation. Nat Cell Biol 19:915–927PubMedCrossRef
51.
Zurück zum Zitat Ninov N, Borius M, Stainier D (2012) Different levels of Notch signaling regulate quiescence, renewal and differentiation in pancreatic endocrine progenitors. Development 139:1557–1567PubMedPubMedCentralCrossRef Ninov N, Borius M, Stainier D (2012) Different levels of Notch signaling regulate quiescence, renewal and differentiation in pancreatic endocrine progenitors. Development 139:1557–1567PubMedPubMedCentralCrossRef
52.
Zurück zum Zitat Stenmark H, Parton R, Steele-Mortimer O, Lütcke A, Gruenberg J, Zerial M (1994) Inhibition of rab5 GTPase activity stimulates membrane fusion in endocytosis. EMBO J 13:1287–1296PubMedPubMedCentralCrossRef Stenmark H, Parton R, Steele-Mortimer O, Lütcke A, Gruenberg J, Zerial M (1994) Inhibition of rab5 GTPase activity stimulates membrane fusion in endocytosis. EMBO J 13:1287–1296PubMedPubMedCentralCrossRef
55.
Zurück zum Zitat Sandri C, Caccavari F, Valdembri D, Camillo C, Veltel S, Santambrogio M, Lanzetti L, Bussolino F, Ivaska J, Serini G (2012) The R-Ras/RIN2/Rab5 complex controls endothelial cell adhesion and morphogenesis via active integrin endocytosis and Rac signaling. Cell Res 22:1479–1501PubMedPubMedCentralCrossRef Sandri C, Caccavari F, Valdembri D, Camillo C, Veltel S, Santambrogio M, Lanzetti L, Bussolino F, Ivaska J, Serini G (2012) The R-Ras/RIN2/Rab5 complex controls endothelial cell adhesion and morphogenesis via active integrin endocytosis and Rac signaling. Cell Res 22:1479–1501PubMedPubMedCentralCrossRef
56.
Zurück zum Zitat Jakobsson L, Franco C, Bentley K, Collins R, Ponsioen B, Aspalter I, Rosewell I, Busse M, Thurston G, Medvinsky A, Schulte-Merker S, Gerhardt H (2010) Endothelial cells dynamically compete for the tip cell position during angiogenic sprouting. Nat Cell Biol 12:943–953PubMedCrossRef Jakobsson L, Franco C, Bentley K, Collins R, Ponsioen B, Aspalter I, Rosewell I, Busse M, Thurston G, Medvinsky A, Schulte-Merker S, Gerhardt H (2010) Endothelial cells dynamically compete for the tip cell position during angiogenic sprouting. Nat Cell Biol 12:943–953PubMedCrossRef
57.
Zurück zum Zitat Lucitti J, Sealock R, Buckley B, Zhang H, Xiao L, Dudley A, Faber J (2016) Variants of Rab GTPase-effector binding protein-2 cause variation in the collateral circulation and severity of stroke. Stroke 47:3022–3031PubMedPubMedCentralCrossRef Lucitti J, Sealock R, Buckley B, Zhang H, Xiao L, Dudley A, Faber J (2016) Variants of Rab GTPase-effector binding protein-2 cause variation in the collateral circulation and severity of stroke. Stroke 47:3022–3031PubMedPubMedCentralCrossRef
58.
Zurück zum Zitat Barrett T, Wilhite SE, Ledoux P, Evangelista C, Kim IF, Tomashevsky M, Marshall KA, Phillippy KH, Sherman PM, Holko M, Yefanov A, Lee H, Zhang N, Robertson CL, Serova N, Davis S, Soboleva A (2013) NCBI GEO: archive for functional genomics data sets—update. Nucleic Acids Res 41:991–995CrossRef Barrett T, Wilhite SE, Ledoux P, Evangelista C, Kim IF, Tomashevsky M, Marshall KA, Phillippy KH, Sherman PM, Holko M, Yefanov A, Lee H, Zhang N, Robertson CL, Serova N, Davis S, Soboleva A (2013) NCBI GEO: archive for functional genomics data sets—update. Nucleic Acids Res 41:991–995CrossRef
59.
Zurück zum Zitat Revet I, Huizenga G, Chan A, Koster J, Volckmann R, van Sluis P, Øra I, Versteeg R, Geerts D (2008) The MSX1 homeobox transcription factor is a downstream target of PHOX2B and activates the Delta-Notch pathway in neuroblastoma. Exp Cell Res 314:707–719PubMedCrossRef Revet I, Huizenga G, Chan A, Koster J, Volckmann R, van Sluis P, Øra I, Versteeg R, Geerts D (2008) The MSX1 homeobox transcription factor is a downstream target of PHOX2B and activates the Delta-Notch pathway in neuroblastoma. Exp Cell Res 314:707–719PubMedCrossRef
60.
Zurück zum Zitat Eglinger J, Karsjens H, Lammert E (2017) Quantitative assessment of angiogenesis and pericyte coverage in human cell-derived vascular sprouts. Inflamm Regen 37:1–9CrossRef Eglinger J, Karsjens H, Lammert E (2017) Quantitative assessment of angiogenesis and pericyte coverage in human cell-derived vascular sprouts. Inflamm Regen 37:1–9CrossRef
61.
Zurück zum Zitat Westerfield M (1993) The Zebrafish book. University of Oregon Press, Eugene Westerfield M (1993) The Zebrafish book. University of Oregon Press, Eugene
62.
Zurück zum Zitat Kimmel C, Ballard W, Kimmel S, Ullmann B, Schilling T (1995) Stages of embryonic development of the Zebrafish. Dev Dyn 203:253–310PubMedCrossRef Kimmel C, Ballard W, Kimmel S, Ullmann B, Schilling T (1995) Stages of embryonic development of the Zebrafish. Dev Dyn 203:253–310PubMedCrossRef
63.
Zurück zum Zitat Chi N, Shaw R, De Val S, Kang G, Jan L, Black B, Stainier D (2008) Foxn4 directly regulates tbx2b expression and atrioventricular canal formation. Genes Dev 22:734–739PubMedPubMedCentralCrossRef Chi N, Shaw R, De Val S, Kang G, Jan L, Black B, Stainier D (2008) Foxn4 directly regulates tbx2b expression and atrioventricular canal formation. Genes Dev 22:734–739PubMedPubMedCentralCrossRef
64.
Zurück zum Zitat Jin S, Beis D, Mitchell T, Chen J, Stainier D (2005) Cellular and molecular analyses of vascular tube and lumen formation in zebrafish. Development 132:5199–5209PubMedCrossRef Jin S, Beis D, Mitchell T, Chen J, Stainier D (2005) Cellular and molecular analyses of vascular tube and lumen formation in zebrafish. Development 132:5199–5209PubMedCrossRef
65.
Zurück zum Zitat Helker C, Schuermann A, Karpanen T, Zeuschner D, Belting H-G, Affolter M, Schulte-Merker S, Herzog W (2013) The zebrafish common cardinal veins develop by a novel mechanism: lumen ensheathment. Development 140:2776–2786PubMedCrossRef Helker C, Schuermann A, Karpanen T, Zeuschner D, Belting H-G, Affolter M, Schulte-Merker S, Herzog W (2013) The zebrafish common cardinal veins develop by a novel mechanism: lumen ensheathment. Development 140:2776–2786PubMedCrossRef
66.
Zurück zum Zitat Hruscha A, Krawitz P, Rechenberg A, Heinrich V, Hecht J, Haass C, Schmid B (2013) Efficient CRISPR/Cas9 genome editing with low off-target effects in zebrafish. Development 140:4982–4987PubMedCrossRef Hruscha A, Krawitz P, Rechenberg A, Heinrich V, Hecht J, Haass C, Schmid B (2013) Efficient CRISPR/Cas9 genome editing with low off-target effects in zebrafish. Development 140:4982–4987PubMedCrossRef
67.
Zurück zum Zitat Wakayama Y, Fukuhara S, Ando K, Matsuda M, Mochizuki N (2015) Cdc42 mediates Bmp-induced sprouting angiogenesis through Fmnl3-driven assembly of endothelial filopodia in zebrafish. Dev Cell 32:109–122PubMedCrossRef Wakayama Y, Fukuhara S, Ando K, Matsuda M, Mochizuki N (2015) Cdc42 mediates Bmp-induced sprouting angiogenesis through Fmnl3-driven assembly of endothelial filopodia in zebrafish. Dev Cell 32:109–122PubMedCrossRef
68.
Zurück zum Zitat Hübner K, Cabochette P, Diéguez-Hurtado R, Wiesner C, Wakayama Y, Grassme KS, Hubert M, Guenther S, Belting HG, Affolter M, Adams RH (2018) Wnt/β-catenin signaling regulates VE-cadherin-mediated anastomosis of brain capillaries by counteracting S1pr1 signaling. Nat Commun 9(1):1–7CrossRef Hübner K, Cabochette P, Diéguez-Hurtado R, Wiesner C, Wakayama Y, Grassme KS, Hubert M, Guenther S, Belting HG, Affolter M, Adams RH (2018) Wnt/β-catenin signaling regulates VE-cadherin-mediated anastomosis of brain capillaries by counteracting S1pr1 signaling. Nat Commun 9(1):1–7CrossRef
69.
Zurück zum Zitat Ritchie M, Phipson B, Wu D, Hu Y, Law C, Shi W, Smyth G (2015) limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res 43:e47PubMedPubMedCentralCrossRef Ritchie M, Phipson B, Wu D, Hu Y, Law C, Shi W, Smyth G (2015) limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res 43:e47PubMedPubMedCentralCrossRef
70.
Zurück zum Zitat Robinson M, McCarthy D, Smyth G (2010) edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics 26:139–140PubMedCrossRef Robinson M, McCarthy D, Smyth G (2010) edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics 26:139–140PubMedCrossRef
Metadaten
Titel
The endosomal RIN2/Rab5C machinery prevents VEGFR2 degradation to control gene expression and tip cell identity during angiogenesis
verfasst von
Lanette Kempers
Yuki Wakayama
Ivo van der Bijl
Charita Furumaya
Iris M. De Cuyper
Aldo Jongejan
Marije Kat
Anne-Marieke D. van Stalborch
Antonius L. van Boxtel
Marvin Hubert
Dirk Geerts
Jaap D. van Buul
Dirk de Korte
Wiebke Herzog
Coert Margadant
Publikationsdatum
13.05.2021
Verlag
Springer Netherlands
Erschienen in
Angiogenesis / Ausgabe 3/2021
Print ISSN: 0969-6970
Elektronische ISSN: 1573-7209
DOI
https://doi.org/10.1007/s10456-021-09788-4

Weitere Artikel der Ausgabe 3/2021

Angiogenesis 3/2021 Zur Ausgabe

Leitlinien kompakt für die Innere Medizin

Mit medbee Pocketcards sicher entscheiden.

Seit 2022 gehört die medbee GmbH zum Springer Medizin Verlag

Echinokokkose medikamentös behandeln oder operieren?

06.05.2024 DCK 2024 Kongressbericht

Die Therapie von Echinokokkosen sollte immer in spezialisierten Zentren erfolgen. Eine symptomlose Echinokokkose kann – egal ob von Hunde- oder Fuchsbandwurm ausgelöst – konservativ erfolgen. Wenn eine Op. nötig ist, kann es sinnvoll sein, vorher Zysten zu leeren und zu desinfizieren. 

Umsetzung der POMGAT-Leitlinie läuft

03.05.2024 DCK 2024 Kongressbericht

Seit November 2023 gibt es evidenzbasierte Empfehlungen zum perioperativen Management bei gastrointestinalen Tumoren (POMGAT) auf S3-Niveau. Vieles wird schon entsprechend der Empfehlungen durchgeführt. Wo es im Alltag noch hapert, zeigt eine Umfrage in einem Klinikverbund.

Proximale Humerusfraktur: Auch 100-Jährige operieren?

01.05.2024 DCK 2024 Kongressbericht

Mit dem demographischen Wandel versorgt auch die Chirurgie immer mehr betagte Menschen. Von Entwicklungen wie Fast-Track können auch ältere Menschen profitieren und bei proximaler Humerusfraktur können selbst manche 100-Jährige noch sicher operiert werden.

Die „Zehn Gebote“ des Endokarditis-Managements

30.04.2024 Endokarditis Leitlinie kompakt

Worauf kommt es beim Management von Personen mit infektiöser Endokarditis an? Eine Kardiologin und ein Kardiologe fassen die zehn wichtigsten Punkte der neuen ESC-Leitlinie zusammen.

Update Innere Medizin

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.