Neuropilin 1 (NRP1) is a receptor for class 3 semaphorins and vascular endothelial growth factor (VEGF) A and is essential for cardiovascular development. Biochemical evidence supports a model for NRP1 function in which VEGF binding induces complex formation between NRP1 and VEGFR2 to enhance endothelial VEGF signalling. However, the relevance of VEGF binding to NRP1 for angiogenesis in vivo has not yet been examined. We therefore generated knock-in mice expressing Nrp1 with a mutation of tyrosine (Y) 297 in the VEGF binding pocket of the NRP1 b1 domain, as this residue was previously shown to be important for high affinity VEGF binding and NRP1-VEGFR2 complex formation. Unexpectedly, this targeting strategy also severely reduced NRP1 expression and therefore generated a NRP1 hypomorph. Despite the loss of VEGF binding and attenuated NRP1 expression, homozygous Nrp1Y297A/Y297A mice were born at normal Mendelian ratios, arguing against NRP1 functioning exclusively as a VEGF164 receptor in embryonic angiogenesis. By overcoming the mid-gestation lethality of full Nrp1-null mice, homozygous Nrp1Y297A/Y297A mice revealed essential roles for NRP1 in postnatal angiogenesis and arteriogenesis in the heart and retina, pathological neovascularisation of the retina and angiogenesis-dependent tumour growth.

NRP1 is a transmembrane receptor for the VEGF165 isoform (VEGF164 in mice) and the neuronal guidance cue SEMA3A, with essential roles in both vascular and neuronal development (reviewed by Pellet-Many et al., 2008; Raimondi and Ruhrberg, 2013). Accordingly, Nrp1-null mice die before birth with severe cardiovascular and neuronal defects (Kitsukawa et al., 1997; Kawasaki et al., 1999). Mice lacking NRP1 in vascular endothelial cells (ECs) also show defective cardiovascular development, whereas mice carrying a mutated extracellular domain that abolishes SEMA3A, not VEGF164, binding show defective nerve, but not blood vessel, patterning (Gu et al., 2003; Vieira et al., 2007). These and other genetic, biochemical and cell biological data support a model in which VEGF165 binding induces complex formation between NRP1 and VEGFR2 (KDR - Mouse Genome Informatics) to enhance VEGFR2 signalling during EC migration in vitro (e.g. Soker et al., 2002; Wang et al., 2003; Evans et al., 2011) and arteriogenesis in vivo (Lanahan et al., 2013).

The extracellular NRP1 a1/a2 and b1/b2 domains are crucial for recognition of SEMA3A, whereas the major VEGF binding site resides in the b1 domain, with some contribution of the b2 domain (Gu et al., 2002). Structural studies of the b1 domain have identified key residues important for ligand interactions in a VEGF165-binding pocket (von Wronski et al., 2006; Vander Kooi et al., 2007; Jarvis et al., 2010). For example, mutation of Y297 or D320 in the b1 domain impairs VEGF165 binding and complex formation between NRP1 and VEGFR2, which reduces VEGF165-stimulated EC migration in vitro (Jarvis et al., 2010; Herzog et al., 2011).

Though NRP1 is essential for developmental angiogenesis, the contribution of the NRP1-VEGF interaction to physiological or pathological angiogenesis is unclear. We therefore generated mice with a Y297A mutation in the b1 VEGF binding site (supplementary material Fig. S1A). Inadvertently, insertion of a Nrp1Y297A mutant cDNA reduced mRNA and protein expression of the mutant allele, generating a novel mouse model with NRP1 hypomorphism. Despite combining loss of VEGF binding with reduced NRP1 expression, Nrp1Y297A/Y297A homozygous mice were mostly viable at birth, without the severe and lethal cardiovascular phenotypes of full or endothelial-specific Nrp1-null mutants. This observation implies that NRP1 has important VEGF-independent roles in angiogenesis which most likely synergise with its known role as a VEGFR2 co-receptor. Moreover, the postnatal viability of Nrp1Y297A/Y297A mice enabled the study of newborn and adult mutants, thereby extending genetic evidence for the established function of NRP1 in cardiovascular development to include essential roles in promoting postnatal and pathological angiogenesis.

Reduced NRP1 expression in Nrp1Y297A/Y297A mice

Previous studies demonstrated that the highly conserved Y297 residue in the NRP1 b1 domain (supplementary material Fig. S1A,B) is essential for VEGF165 binding to human NRP1 (Jarvis et al., 2010; Herzog et al., 2011). The Y297A mutation also inhibited high affinity binding of 125I-VEGF165 to murine NRP1 in vitro (supplementary material Fig. S1C). Heterozygous mice expressing a mutated Nrp1Y297A cDNA from exon 2 of the endogenous Nrp1 locus were generated (Fig. 1A) and interbred to produce homozygous Nrp1Y297A/Y297A offspring (Fig. 1B-D). Unexpectedly, immunoblotting revealed a severe reduction of NRP1 in Nrp1Y297A/Y297A and a moderate reduction in Nrp1Y297A/+ mice compared with wild-type (WT) littermates (Fig. 1E,F). Immunohistochemistry confirmed reduced NRP1 expression in Nrp1Y297A/Y297A tissues (supplementary material Fig. S2A,B). Reduced NRP1 protein levels correlated with decreased Nrp1 mRNA expression in all Nrp1Y297A/Y297A tissues examined (supplementary material Fig. S2C). Impaired expression of the Nrp1Y297A allele may be due to the disruption of intronic regulatory elements deleted by insertion of the mutant cDNA into the Nrp1 locus. Allele-specific quantitative PCR (qPCR) of heterozygous mice demonstrated significantly reduced expression of the mutant allele compared with the WT allele in all tissues examined (Fig. 1G). Mice expressing the Nrp1Y297A allele are therefore hypomorphic for NRP1 with markedly reduced NRP1 expression as well as reduced VEGF164 binding and represent a novel model in which to examine NRP1 roles in angiogenesis in vivo.

Fig. 1.

Nrp1Y297A/Y297A mice have reduced NRP1 expression, survival, body weight and heart vascularisation. (A) Mouse Nrp1 cDNA with a Y297A substitution was used to target exon 2 of the endogenous Nrp1 locus. (B) Southern blot of AflII-digested genomic DNA from WT and gene-targeted mice after removal of the neomycin cassette. (C) PCR analysis of WT and mutant alleles. (D) cDNA sequencing of WT and homozygous mice. (E,F) NRP1 levels in immunoblots were quantified by densitometry relative to GAPDH or β-actin (mean ± s.d.; n=6 each). (G) Nrp1 mRNA expression from WT (black bars) and Nrp1Y297A (white bars) alleles in Nrp1+/Y297A mice was determined by TaqMan qPCR (mean ± s.d.; n=6). (H) Survival of WT and Nrp1Y297A/Y297A littermates during the first 6 weeks of postnatal life (n≥27). (I) Body weights of WT and Nrp1Y297A/Y297A littermates at P7 (n=5-11; mean ± s.d.). (J) Body weight of WT and Nrp1Y297A/Y297A males between 6 and 18 weeks (n=3-10 animals per group and genotype; mean ± s.d.). (K) Appearance of 6-week-old WT and Nrp1Y297A/Y297A littermates. (L) Post mortem analysis of a suddenly deceased P7 Nrp1Y297A/Y297A pup. Arrowhead indicates blood-filled lungs. (M-P) Immunostaining for endomucin (green), SMA (red) and DAPI (blue) of P7 heart cryosections from WT and Nrp1Y297A/Y297A littermates. (M′,O′) Single SMA channels of panels M,O. (N,P) High magnification image of insets in panels M,O. Arrowheads indicate SMA+ endomucin-coronary arteries; open arrowheads indicate SMA+ endomucin+ coronary vein; arrows indicate endocardium; n=3 each. *P<0.05, **P<0.01, ***P<0.001 Nrp1Y297A/Y297A or Nrp1+/Y297A versus WT. Scale bar: 100 μm (M-P).

Fig. 1.

Nrp1Y297A/Y297A mice have reduced NRP1 expression, survival, body weight and heart vascularisation. (A) Mouse Nrp1 cDNA with a Y297A substitution was used to target exon 2 of the endogenous Nrp1 locus. (B) Southern blot of AflII-digested genomic DNA from WT and gene-targeted mice after removal of the neomycin cassette. (C) PCR analysis of WT and mutant alleles. (D) cDNA sequencing of WT and homozygous mice. (E,F) NRP1 levels in immunoblots were quantified by densitometry relative to GAPDH or β-actin (mean ± s.d.; n=6 each). (G) Nrp1 mRNA expression from WT (black bars) and Nrp1Y297A (white bars) alleles in Nrp1+/Y297A mice was determined by TaqMan qPCR (mean ± s.d.; n=6). (H) Survival of WT and Nrp1Y297A/Y297A littermates during the first 6 weeks of postnatal life (n≥27). (I) Body weights of WT and Nrp1Y297A/Y297A littermates at P7 (n=5-11; mean ± s.d.). (J) Body weight of WT and Nrp1Y297A/Y297A males between 6 and 18 weeks (n=3-10 animals per group and genotype; mean ± s.d.). (K) Appearance of 6-week-old WT and Nrp1Y297A/Y297A littermates. (L) Post mortem analysis of a suddenly deceased P7 Nrp1Y297A/Y297A pup. Arrowhead indicates blood-filled lungs. (M-P) Immunostaining for endomucin (green), SMA (red) and DAPI (blue) of P7 heart cryosections from WT and Nrp1Y297A/Y297A littermates. (M′,O′) Single SMA channels of panels M,O. (N,P) High magnification image of insets in panels M,O. Arrowheads indicate SMA+ endomucin-coronary arteries; open arrowheads indicate SMA+ endomucin+ coronary vein; arrows indicate endocardium; n=3 each. *P<0.05, **P<0.01, ***P<0.001 Nrp1Y297A/Y297A or Nrp1+/Y297A versus WT. Scale bar: 100 μm (M-P).

Increased mortality and impaired growth in postnatal Nrp1Y297A/Y297A mice

Genotyping of offspring from crosses between heterozygous mice yielded the expected Mendelian ratio for WT, heterozygous and homozygous mice before birth, and only a small, non-significant reduction in the number of Nrp1Y297A/Y297A offspring on postnatal day (P) 0 (supplementary material Table S2), indicating that residual NRP1 expression from the Nrp1Y297A allele prevents the mid-gestation lethality of full NRP1 knockouts. By contrast, homozygous mutants were under-represented in litters at P21, and the number of adult Nrp1Y297A/Y297A mice (>8 weeks) was ∼50% lower than expected (supplementary material Table S2). Follow-up from birth to 6 weeks of age showed increased mortality of Nrp1Y297A/Y297A mice predominantly within the first 2 weeks of postnatal life (Fig. 1H). Nrp1Y297A/Y297A mice were also significantly smaller than WT littermates at P7 (Fig. 1I) and at all ages examined between 6 and 18 weeks (Fig. 1J,K). Heterozygous mice had normal weights at P7, but showed a small, significant size reduction at 8, 10, 12 and 14 weeks (data not shown). These results indicate that low NRP1 expression combined with inhibition of VEGF164 binding causes postnatal mortality and reduces body weight in surviving mice.

Defective myocardial vascularisation in postnatal Nrp1Y297A/Y297A mice

Post mortem analysis of mutants that had died on P7 revealed blood-filled lungs (Fig. 1L), indicative of congestive heart failure. Increased mortality and reduced body size are also seen in Vegfa120/120 mutants that die perinatally owing to ischemic cardiomyopathy caused by impaired myocardial vascularisation (Carmeliet et al., 1999). Similarly, P7 Nrp1Y297A/Y297A hearts contained fewer coronary arteries and capillaries than control hearts (Fig. 1M-P). Moreover, capillaries in mutants appeared abnormal and were sparse, particularly in the subendomyocardium, consistent with reports that angiogenesis occurs in an epicardial-to-endocardial gradient (Tomanek, 1996). The similar vascular defects in Nrp1Y297A/Y297A and Vegfa120/120 myocardium, combined with a predominance of NRP-binding VEGF isoforms relative to VEGF120 in this tissue (Ng et al., 2001), suggest that the perinatal vascularisation of myocardium depends on VEGF binding to NRP1.

Embryonic angiogenesis is mildly impaired in Nrp1Y297A/Y297A mutants

We next investigated the effect of the Y297A mutation on VEGF binding to NRP1 and on vascular development in the hindbrain on embryonic days (E) 11.5 and E12.5, when blood vessels sprout and fuse to form the subventricular vascular plexus (SVP) (Fantin et al., 2010). Although Nrp2, Vegfr1 (Flt1 - Mouse Genome Informatics) and Vegfr2 mRNA levels were unaffected, Nrp1 transcripts were reduced by 50% in Nrp1Y297A/Y297A compared with WT hindbrains (Fig. 2A). In agreement, immunostaining showed reduced NRP1 protein levels in the mutants, both in axons and vessels (Fig. 2B,C). Alkaline phosphatase (AP)-conjugated VEGF165 ligand binding to NRP1-positive pial axons was abolished in Nrp1Y297A/Y297A mutant, similar to Nrp1-null, hindbrains (Fig. 2D). By contrast, AP-VEGF165 binding to vessels was unaffected (Fig. 2D), consistent with high VEGFR2 expression by hindbrain ECs, but not axons (Lanahan et al., 2013). AP-SEMA3A binding to NRP1-expressing axons in Nrp1Y297A/Y297A mice was unperturbed (Fig. 2E). However, AP-SEMA3A binding to SVP vessels was impaired, even though binding to radial vessels on the pial side was unaffected (Fig. 2E), perhaps due to the differential effect of the Nrp1 mutation on overall NRP1 protein levels in the two brain regions.

Fig. 2.

Mild angiogenesis defects in Nrp1Y297A/Y297A hindbrains. (A-C) VEGF receptor expression in Nrp1Y297A/Y297A hindbrains. qPCR analysis for Nrp1, Nrp2, Vegfr1 and Vegfr2 relative to Actb (A) (n=3; mean ± s.d.; **P<0.01, ns, not significant) and whole-mount immunofluorescence staining for NRP1 and IB4 (B) in E11.5 hindbrains. (C) Single NRP1 channels are shown in greyscale and heat-map (red, blue and green indicate high, low and medium pixel intensity, respectively). Arrowheads indicate axons. (D,E) AP-VEGF165 (D) and AP-SEMA3A (E) binding in E12.5 Nrp1Y297A/Y297A hindbrains; arrowheads indicate axons; asterisks highlight absent ligand binding to axons; open triangles indicate absent SEMA3A vessel binding. (F-I) Whole-mount PECAM immunostaining and quantification of vessels in E12.5 hindbrains, imaged (F) and quantified (G) on the pial side or imaged (H) and quantified (I) on the SVP side; the arrow indicates a vascular tuft; n≥4; mean ± s.d.; ***P<0.001. Scale bar: in C, 100 μm (B-F,H).

Fig. 2.

Mild angiogenesis defects in Nrp1Y297A/Y297A hindbrains. (A-C) VEGF receptor expression in Nrp1Y297A/Y297A hindbrains. qPCR analysis for Nrp1, Nrp2, Vegfr1 and Vegfr2 relative to Actb (A) (n=3; mean ± s.d.; **P<0.01, ns, not significant) and whole-mount immunofluorescence staining for NRP1 and IB4 (B) in E11.5 hindbrains. (C) Single NRP1 channels are shown in greyscale and heat-map (red, blue and green indicate high, low and medium pixel intensity, respectively). Arrowheads indicate axons. (D,E) AP-VEGF165 (D) and AP-SEMA3A (E) binding in E12.5 Nrp1Y297A/Y297A hindbrains; arrowheads indicate axons; asterisks highlight absent ligand binding to axons; open triangles indicate absent SEMA3A vessel binding. (F-I) Whole-mount PECAM immunostaining and quantification of vessels in E12.5 hindbrains, imaged (F) and quantified (G) on the pial side or imaged (H) and quantified (I) on the SVP side; the arrow indicates a vascular tuft; n≥4; mean ± s.d.; ***P<0.001. Scale bar: in C, 100 μm (B-F,H).

E12.5 Nrp1Y297A/Y297A hindbrains exhibited a similar number of radial vessels compared with WT littermates (Fig. 2F,G), but significantly reduced vessel branching in the SVP (Fig. 2H,I), as observed for Nrp1-null mice (Gerhardt et al., 2004; Fantin et al., 2013a). Yet, the effect of the Nrp1Y297A/Y297A mutation on SVP angiogenesis was less severe than that of complete or endothelial-specific NRP1 loss, which causes a catastrophic failure of neighbouring vessels to interconnect, giving rise to large vascular tufts at vessel termini (Gerhardt et al., 2004; Fantin et al., 2013a). Importantly, the mild vascular defects in Nrp1Y297A/Y297A hindbrains do not phenocopy Vegfa120/120 hindbrains expressing VEGF120, but lacking the NRP1-binding VEGF isoforms; specifically, Vegfa120/120 hindbrains contain fewer pial vessels as well as poorly branched and enlarged vessels throughout the brain, and these defects can be attributed to perturbed VEGF gradients (Ruhrberg et al., 2002). As in the hindbrain, vessel density was only mildly affected in head and trunk of E9.5 Nrp1Y297A/Y297A embryos, and the formation of the pharyngeal arch arteries and the intersomitic vessels appeared to be unaffected (supplementary material Fig. S3). The mild vascular phenotype of Nrp1Y297A/Y297A embryos compared with the severe phenotype of Nrp1-null and Vegfa120/120 mice argues against a major role for VEGF164-signalling through NRP1 in embryonic angiogenesis.

Retinal angiogenesis and arteriogenesis are defective in Nrp1Y297A/Y297A mice

The impact of the Y297A mutation on postnatal vascular development was further investigated in the retina, where angiogenesis begins at P0 and is essentially complete by P21 (Fruttiger, 2007). Examination of the P7 retinal vasculature by wholemount immunofluorescence for the vascular marker collagen IV revealed a significant reduction in the extension of the primary vascular plexus towards the retinal margin in Nrp1Y297A/Y297A compared with WT retinas (Fig. 3A,B). The number of arteries and veins was also reduced in Nrp1Y297A/Y297A retinas (Fig. 3A,C; data not shown), and mutants displayed significantly reduced vessel coverage by vascular smooth muscle cells (VSMCs) positive for smooth muscle α-actin (SMA) (Fig. 3A,D,E). Similar to the hindbrain, the Nrp1Y297A mutation greatly reduced NRP1 expression in Nrp1Y297A/Y297A compared with WT retinas (Fig. 3F).

Fig. 3.

Defective retinal angiogenesis in Nrp1Y297A/Y297A mice. (A-M) Immunostaining of retinal flatmounts and quantification of vascular patterning at P7 (A-E), P10 (F) and P21 (G-M). WT and Nrp1Y297A/Y297A littermates were immunolabelled for collagen IV and SMA (A,G,J; single SMA channel in greyscale), NRP1 (F; single NRP1 channel in greyscale and as pixel intensity heat-map) or IB4 and SMA with NG2 (K) or collagen IV (L,M). NRP1-positive cells at the mutant vascular front in F resemble macrophages normally observed at earlier stages of vascular development and therefore indicate delayed vascularisation. The area indicated with a dotted square in K is shown in higher magnification on the right. (L,M) Examples of defective main vessel segregation in Nrp1Y297A/Y297A mutants: multiple artery/vein crossings (L) and arterial anastomosis (M). Quantification of vascular extension (B), artery number (C,H), extension of SMA+ arteries (D) and arterial coverage by SMA+ VSMC (E,I); n≥3; mean ± s.d.; ***P<0.001. Scale bars: 1 mm (A,G); 200 μm (F,K-M); 25 μm (J).

Fig. 3.

Defective retinal angiogenesis in Nrp1Y297A/Y297A mice. (A-M) Immunostaining of retinal flatmounts and quantification of vascular patterning at P7 (A-E), P10 (F) and P21 (G-M). WT and Nrp1Y297A/Y297A littermates were immunolabelled for collagen IV and SMA (A,G,J; single SMA channel in greyscale), NRP1 (F; single NRP1 channel in greyscale and as pixel intensity heat-map) or IB4 and SMA with NG2 (K) or collagen IV (L,M). NRP1-positive cells at the mutant vascular front in F resemble macrophages normally observed at earlier stages of vascular development and therefore indicate delayed vascularisation. The area indicated with a dotted square in K is shown in higher magnification on the right. (L,M) Examples of defective main vessel segregation in Nrp1Y297A/Y297A mutants: multiple artery/vein crossings (L) and arterial anastomosis (M). Quantification of vascular extension (B), artery number (C,H), extension of SMA+ arteries (D) and arterial coverage by SMA+ VSMC (E,I); n≥3; mean ± s.d.; ***P<0.001. Scale bars: 1 mm (A,G); 200 μm (F,K-M); 25 μm (J).

By P21, delayed angiogenesis in the primary vascular plexus had largely recovered in homozygous mutants (Fig. 3G), but delayed formation of the deep and intermediate plexi was indicated by reduced vascular density (supplementary material Fig. S4). The number of arteries and their VSMC coverage remained significantly reduced (Fig. 3G-J). Moreover, incomplete vascularisation was observed in parts of the peripheral retinas, where vessels terminated in endothelial tufts covered with SMA-positive VSMCs and NG2 (CSPG4 - Mouse Genome Informatics)-positive pericytes (Fig. 3K). As NRP1 is not essential for endothelial cell proliferation (Jones et al., 2008), but is important for migration (Wang et al., 2003; Pan et al., 2007; Evans et al., 2011), tuft formation is likely to be due to abnormal endothelial cell sprouting. Nrp1Y297A/Y297A retinas also contained an increased number of arteriovenous crossings (Fig. 3L,M), similar to mice lacking the NRP1 cytoplasmic domain (Fantin et al., 2011) or with haploinsufficient expression of VEGF in the neural lineage (Haigh et al., 2003).

We cannot presently resolve whether low NRP1 expression levels and/or inhibition of VEGF164 binding to NRP1 are primarily responsible for the observed retinal vascular defects in Nrp1Y297A/Y297A mice. However, their similarity to defects in Vegfa120/120 retinas (Stalmans et al., 2002), combined with similar vascular defects in the P7 myocardium (Fig. 1M-P), suggests that NRP1-mediated VEGF-signalling is more important for perinatal than embryonic vascular development.

Reduced angiogenesis in Nrp1Y297A/Y297A mice with oxygen-induced retinopathy

The postnatal viability of Nrp1Y297A/Y297A mice allowed us to examine the impact of Nrp1 hypomorphism and defective VEGF binding on pathological angiogenesis. We used a model of oxygen-induced retinopathy, in which sequential exposure of neonatal mice to hyperoxia and normoxia causes obliteration of central retinal capillaries, followed by hypoxia-induced VEGF upregulation and abnormal neovascularisation (Smith et al., 1994). Vaso-obliteration was slightly reduced, consistent with reduced vascular dropout after hyperoxia or increased revascularisation of avascular areas after return to normoxia, whereas pathological neovascularisation was strongly and significantly decreased in Nrp1Y297A/Y297A compared with WT littermates (Fig. 4A-D). These observations show that NRP1 contributes substantially to ocular neovascularisation.

Fig. 4.

Reduced pathological angiogenesis and tumourigenesis in Nrp1Y297A/Y297A mice. (A-D) Reduced angiogenesis in Nrp1Y297A/Y297A mice with oxygen-induced retinopathy. (A-B′) P17 retinal flatmounts from P17 Nrp1+/+ (A-A′) and Nrp1Y297A/Y297A (B-B′) littermates were immunolabelled for collagen IV. Scale bar: 1 mm. Higher magnifications are shown in A′,B′, with pseudocoloured vaso-obliterated (VO, yellow) and neovascular (NV, red) areas in A′,B′. (C,D) Quantification of vaso-obliteration (C) and neovascularisation (D) in mutants relative to WT littermates. (E-H) Reduced VEGF-induced vessel sprouting in Nrp1Y297A/Y297A aortic rings. Aortic rings from adult WT (E), Nrp1+/Y297A (F) and Nrp1Y297A/Y297A (G) littermates were treated with 0.5% FBS ± 50 ng/ml VEGF164 (E-G) or 10% FBS and the total area (H) of angiogenic sprouts quantified after 6 days; n≥4; mean ± s.e.m. Scale bar: 250 μm. (I-L) Reduced B16-F1 melanoma growth in Nrp1+/Y297A and Nrp1Y297A/Y297A mice. Representative tumours (I-K) and tumour weights (L) are shown; n≥3; mean ± s.d.; *P<0.05, **P<0.01.

Fig. 4.

Reduced pathological angiogenesis and tumourigenesis in Nrp1Y297A/Y297A mice. (A-D) Reduced angiogenesis in Nrp1Y297A/Y297A mice with oxygen-induced retinopathy. (A-B′) P17 retinal flatmounts from P17 Nrp1+/+ (A-A′) and Nrp1Y297A/Y297A (B-B′) littermates were immunolabelled for collagen IV. Scale bar: 1 mm. Higher magnifications are shown in A′,B′, with pseudocoloured vaso-obliterated (VO, yellow) and neovascular (NV, red) areas in A′,B′. (C,D) Quantification of vaso-obliteration (C) and neovascularisation (D) in mutants relative to WT littermates. (E-H) Reduced VEGF-induced vessel sprouting in Nrp1Y297A/Y297A aortic rings. Aortic rings from adult WT (E), Nrp1+/Y297A (F) and Nrp1Y297A/Y297A (G) littermates were treated with 0.5% FBS ± 50 ng/ml VEGF164 (E-G) or 10% FBS and the total area (H) of angiogenic sprouts quantified after 6 days; n≥4; mean ± s.e.m. Scale bar: 250 μm. (I-L) Reduced B16-F1 melanoma growth in Nrp1+/Y297A and Nrp1Y297A/Y297A mice. Representative tumours (I-K) and tumour weights (L) are shown; n≥3; mean ± s.d.; *P<0.05, **P<0.01.

Reduced VEGF-induced angiogenesis and tumourigenesis in Nrp1Y297A/Y297A mice

VEGF stimulation of pathological angiogenesis in Nrp1Y297A/Y297A mice was assessed in the aortic ring assay, in which adult aortic explants undergo angiogenic sprouting upon growth factor stimulation (Baker et al., 2011). VEGF164 increased vessel sprouting from WT aortic rings, but this response was suppressed by the Nrp1Y297A mutation (Fig. 4E-H). By contrast, serum stimulated sprouting from WT and mutant aortic rings similarly (Fig. 4H). We next examined the effect of the Y297A mutation on angiogenesis-dependent growth of syngeneic B16-F1 murine melanomas (Franco et al., 2002). These tumours produce high levels of VEGF, and their growth is blocked by VEGFR2 inhibition, demonstrating dependence on VEGF-driven angiogenesis (Prewett et al., 1999). Whereas tumours in WT mice grew rapidly, tumours were significantly smaller in age- and sex-matched Nrp1+/Y297A littermates and essentially absent in homozygotes at two weeks (>90% inhibition in homozygous mutants; Fig. 4I-L). Collectively, these genetic data show that NRP1 plays an important role in pathological angiogenesis driven by VEGF. Pan et al. (Pan et al., 2007) demonstrated that function-blocking antibodies targeting VEGF binding to NRP1 reduce human non-small cell lung carcinoma xenograft tumour size in mice by 37%. The greater effect of the Y297A mutation compared with blocking antibodies (>90% versus 37%) may reflect use of different tumour models and/or an additional effect of reducing NRP1 levels compared with inhibiting only VEGF binding to NRP1, perhaps due to the contribution of ligands other than VEGF to tumour angiogenesis.

Conclusions

The consensus model for the endothelial role of NRP1 is that it functions as a receptor for VEGF165 and complexes with VEGFR2 to generate a holoreceptor that optimally transduces angiogenic signals. However, this model has hitherto been based almost entirely on biochemical and cell culture studies, combined with the analysis of a mouse model that abolishes all endothelial NRP1 functions. The severe reduction in NRP1 expression caused by the Nrp1Y297A allele did not permit definitive distinction of VEGF binding from other NRP1 functions requiring normal expression. However, the mild embryonic vascular phenotypes combined with the extended viability of Nrp1Y297A/Y297A relative to full and endothelial-specific NRP1 knockout mice clearly demonstrates for the first time that VEGF binding to NRP1 is not essential for embryonic angiogenesis and, therefore, that NRP1 functions in embryonic ECs predominantly in a VEGF-independent role. Because embryonic vascular functions of NRP1 do not involve semaphorin binding (Gu et al., 2003; Vieira et al., 2007), future studies will be needed to define the precise mechanism of NRP1 function in angiogenesis, taking into account recent reports of NRP1 as a modulator of PDGF, TGFβ and integrin signalling (Valdembri et al., 2009; Cao et al., 2010; Evans et al., 2011; Pellet-Many et al., 2011; Yaqoob et al., 2012).

Mice

Animal experiments were conducted in accordance with the Animal Care and Ethics Guidelines of University College London (UCL) and the United Kingdom Home Office Animals (Scientific Procedures) Act of 1986. Site-directed mutagenesis of mouse Y297 was performed as described (Herzog et al., 2011). The start codon and coding region of exon 2 of murine WT Nrp1 were replaced by homologous recombination in 129Sv/Pas embryonic stem cells (ESCs) with murine cDNA containing Nrp1Y297A, targeted ESCs were injected into C57/BL6 blastocysts and resulting male F1 chimeras were bred with C57BL/6 females expressing CRE to generate heterozygous Nrp1+/Y297A mice lacking a neo cassette (Genoway; Lyon, France; Fig. 1). Genotyping and sequencing of genomic or cDNA from tail biopsies were performed with the primers listed in supplementary material Table S1.

Expression studies

mRNA levels were measured by qPCR using cylophilin as a normaliser and quantified with qStandard software (UCL). Protein expression in tissue lysates was determined by immunoblotting using rabbit anti-NRP1 (1:500; stock code 3725, Cell Signaling) and horseradish peroxidase-conjugated secondary antibody (1:10,000; stock code sc-2030, Santa Cruz), or immunostaining with goat anti-NRP1 (1:50; stock code AF566, R&D Systems) and biotinylated secondary antibody (1:200; stock code E0466, Dako).

Assays

Cell culture, VEGF-binding assay, wholemount immunofluorescence, alkaline phosphatase (AP)-fusion protein binding assay, oxygen-induced retinopathy, aortic ring assays and tumour cell injections were all performed and quantified according to published procedures (Franco et al., 2002; Jia et al., 2006; Connor et al., 2009; Herzog et al., 2011; Baker et al., 2011; Fantin et al., 2013b).

Statistics

Results are presented as means ± s.d. or s.e.m., as specified in each figure legend. Statistical significance of differences between samples was determined by a two-tailed Student’s t-test; more than two datasets or grouped datasets were compared, respectively, by one-way or two-way analysis of variance followed by post-hoc tests. Genotype distribution was analysed using the χ2 test, survival ratios using the Log-rank (Mantel-Cox) test. Statistical analyses were performed with Prism 5 (GraphPad Software); P<0.05 was considered significant.

We thank the staff of the Biological Resources Units at the UCL Institute of Ophthalmology and the Wolfson Institute for Biomedical Research for help with mouse husbandry. We are grateful to the Imaging Facility at the UCL Institute of Ophthalmology for maintenance of the confocal microscopes.

Author contributions

A.F., B.H., C.R. and I.Z designed the research, performed the research, analysed the data and wrote the paper. L.D. performed research. A.P., M.M. and M.Y. performed research and analysed the data.

Funding

This work was supported by a British Heart Foundation (BHF) programme grant [RG/06/003 to I.Z., M.M. and M.Y.]; a BHF project grant [PG/10/86/28622 to C.R. and A.F.]; a BHF PhD studentship [FS/10/54/28680 to A.P.]; a Wellcome Trust New Investigator Award [095623/2/11/2 to C.R., A.F. and L.D.]; and Ark Therapeutics Ltd (B.H.). Deposited in PMC for immediate release.

Baker
M.
,
Robinson
S. D.
,
Lechertier
T.
,
Barber
P. R.
,
Tavora
B.
,
D’Amico
G.
,
Jones
D. T.
,
Vojnovic
B.
,
Hodivala-Dilke
K.
(
2011
).
Use of the mouse aortic ring assay to study angiogenesis.
Nat. Protoc.
7
,
89
104
.
Cao
Y.
,
Szabolcs
A.
,
Dutta
S. K.
,
Yaqoob
U.
,
Jagavelu
K.
,
Wang
L.
,
Leof
E. B.
,
Urrutia
R. A.
,
Shah
V. H.
,
Mukhopadhyay
D.
(
2010
).
Neuropilin-1 mediates divergent R-Smad signaling and the myofibroblast phenotype.
J. Biol. Chem.
285
,
31840
31848
.
Carmeliet
P.
,
Ng
Y. S.
,
Nuyens
D.
,
Theilmeier
G.
,
Brusselmans
K.
,
Cornelissen
I.
,
Ehler
E.
,
Kakkar
V. V.
,
Stalmans
I.
,
Mattot
V.
, et al. 
. (
1999
).
Impaired myocardial angiogenesis and ischemic cardiomyopathy in mice lacking the vascular endothelial growth factor isoforms VEGF164 and VEGF188.
Nat. Med.
5
,
495
502
.
Connor
K. M.
,
Krah
N. M.
,
Dennison
R. J.
,
Aderman
C. M.
,
Chen
J.
,
Guerin
K. I.
,
Sapieha
P.
,
Stahl
A.
,
Willett
K. L.
,
Smith
L. E.
(
2009
).
Quantification of oxygen-induced retinopathy in the mouse: a model of vessel loss, vessel regrowth and pathological angiogenesis.
Nat. Protoc.
4
,
1565
1573
.
Evans
I. M.
,
Yamaji
M.
,
Britton
G.
,
Pellet-Many
C.
,
Lockie
C.
,
Zachary
I. C.
,
Frankel
P.
(
2011
).
Neuropilin-1 signaling through p130Cas tyrosine phosphorylation is essential for growth factor-dependent migration of glioma and endothelial cells.
Mol. Cell. Biol.
31
,
1174
1185
.
Fantin
A.
,
Vieira
J. M.
,
Gestri
G.
,
Denti
L.
,
Schwarz
Q.
,
Prykhozhij
S.
,
Peri
F.
,
Wilson
S. W.
,
Ruhrberg
C.
(
2010
).
Tissue macrophages act as cellular chaperones for vascular anastomosis downstream of VEGF-mediated endothelial tip cell induction.
Blood
116
,
829
840
.
Fantin
A.
,
Schwarz
Q.
,
Davidson
K.
,
Normando
E. M.
,
Denti
L.
,
Ruhrberg
C.
(
2011
).
The cytoplasmic domain of neuropilin 1 is dispensable for angiogenesis, but promotes the spatial separation of retinal arteries and veins.
Development
138
,
4185
4191
.
Fantin
A.
,
Vieira
J. M.
,
Plein
A.
,
Denti
L.
,
Fruttiger
M.
,
Pollard
J. W.
,
Ruhrberg
C.
(
2013a
).
NRP1 acts cell autonomously in endothelium to promote tip cell function during sprouting angiogenesis.
Blood
121
,
2352
2362
.
Fantin
A.
,
Vieira
J. M.
,
Plein
A.
,
Maden
C. H.
,
Ruhrberg
C.
(
2013b
).
The embryonic mouse hindbrain as a qualitative and quantitative model for studying the molecular and cellular mechanisms of angiogenesis.
Nat. Protoc.
8
,
418
429
.
Franco
S.
,
Segura
I.
,
Riese
H. H.
,
Blasco
M. A.
(
2002
).
Decreased B16F10 melanoma growth and impaired vascularization in telomerase-deficient mice with critically short telomeres.
Cancer Res.
62
,
552
559
.
Fruttiger
M.
(
2007
).
Development of the retinal vasculature.
Angiogenesis
10
,
77
88
.
Gerhardt
H.
,
Ruhrberg
C.
,
Abramsson
A.
,
Fujisawa
H.
,
Shima
D.
,
Betsholtz
C.
(
2004
).
Neuropilin-1 is required for endothelial tip cell guidance in the developing central nervous system.
Dev. Dyn.
231
,
503
509
.
Gu
C.
,
Limberg
B. J.
,
Whitaker
G. B.
,
Perman
B.
,
Leahy
D. J.
,
Rosenbaum
J. S.
,
Ginty
D. D.
,
Kolodkin
A. L.
(
2002
).
Characterization of neuropilin-1 structural features that confer binding to semaphorin 3A and vascular endothelial growth factor 165.
J. Biol. Chem.
277
,
18069
18076
.
Gu
C.
,
Rodriguez
E. R.
,
Reimert
D. V.
,
Shu
T.
,
Fritzsch
B.
,
Richards
L. J.
,
Kolodkin
A. L.
,
Ginty
D. D.
(
2003
).
Neuropilin-1 conveys semaphorin and VEGF signaling during neural and cardiovascular development.
Dev. Cell
5
,
45
57
.
Haigh
J. J.
,
Morelli
P. I.
,
Gerhardt
H.
,
Haigh
K.
,
Tsien
J.
,
Damert
A.
,
Miquerol
L.
,
Muhlner
U.
,
Klein
R.
,
Ferrara
N.
, et al. 
. (
2003
).
Cortical and retinal defects caused by dosage-dependent reductions in VEGF-A paracrine signaling.
Dev. Biol.
262
,
225
241
.
Herzog
B.
,
Pellet-Many
C.
,
Britton
G.
,
Hartzoulakis
B.
,
Zachary
I. C.
(
2011
).
VEGF binding to NRP1 is essential for VEGF stimulation of endothelial cell migration, complex formation between NRP1 and VEGFR2, and signaling via FAK Tyr407 phosphorylation.
Mol. Biol. Cell
22
,
2766
2776
.
Jarvis
A.
,
Allerston
C. K.
,
Jia
H.
,
Herzog
B.
,
Garza-Garcia
A.
,
Winfield
N.
,
Ellard
K.
,
Aqil
R.
,
Lynch
R.
,
Chapman
C.
, et al. 
. (
2010
).
Small molecule inhibitors of the neuropilin-1 vascular endothelial growth factor A (VEGF-A) interaction.
J. Med. Chem.
53
,
2215
2226
.
Jia
H.
,
Bagherzadeh
A.
,
Hartzoulakis
B.
,
Jarvis
A.
,
Löhr
M.
,
Shaikh
S.
,
Aqil
R.
,
Cheng
L.
,
Tickner
M.
,
Esposito
D.
, et al. 
. (
2006
).
Characterization of a bicyclic peptide neuropilin-1 (NP-1) antagonist (EG3287) reveals importance of vascular endothelial growth factor exon 8 for NP-1 binding and role of NP-1 in KDR signaling.
J. Biol. Chem.
281
,
13493
13502
.
Jones
E. A.
,
Yuan
L.
,
Breant
C.
,
Watts
R. J.
,
Eichmann
A.
(
2008
).
Separating genetic and hemodynamic defects in neuropilin 1 knockout embryos.
Development
135
,
2479
2488
.
Kawasaki
T.
,
Kitsukawa
T.
,
Bekku
Y.
,
Matsuda
Y.
,
Sanbo
M.
,
Yagi
T.
,
Fujisawa
H.
(
1999
).
A requirement for neuropilin-1 in embryonic vessel formation.
Development
126
,
4895
4902
.
Kitsukawa
T.
,
Shimizu
M.
,
Sanbo
M.
,
Hirata
T.
,
Taniguchi
M.
,
Bekku
Y.
,
Yagi
T.
,
Fujisawa
H.
(
1997
).
Neuropilin-semaphorin III/D-mediated chemorepulsive signals play a crucial role in peripheral nerve projection in mice.
Neuron
19
,
995
1005
.
Lanahan
A.
,
Zhang
X.
,
Fantin
A.
,
Zhuang
Z.
,
Rivera-Molina
F.
,
Speichinger
K.
,
Prahst
C.
,
Zhang
J.
,
Wang
Y.
,
Davis
G.
, et al. 
. (
2013
).
The neuropilin 1 cytoplasmic domain is required for VEGF-A-dependent arteriogenesis.
Dev. Cell
25
,
156
168
.
Ng
Y. S.
,
Rohan
R.
,
Sunday
M. E.
,
Demello
D. E.
,
D’Amore
P. A.
(
2001
).
Differential expression of VEGF isoforms in mouse during development and in the adult.
Dev. Dyn.
220
,
112
121
.
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
.
Pellet-Many
C.
,
Frankel
P.
,
Jia
H.
,
Zachary
I.
(
2008
).
Neuropilins: structure, function and role in disease.
Biochem. J.
411
,
211
226
.
Pellet-Many
C.
,
Frankel
P.
,
Evans
I. M.
,
Herzog
B.
,
Jünemann-Ramírez
M.
,
Zachary
I. C.
(
2011
).
Neuropilin-1 mediates PDGF stimulation of vascular smooth muscle cell migration and signalling via p130Cas.
Biochem. J.
435
,
609
618
.
Prewett
M.
,
Huber
J.
,
Li
Y.
,
Santiago
A.
,
O’Connor
W.
,
King
K.
,
Overholser
J.
,
Hooper
A.
,
Pytowski
B.
,
Witte
L.
, et al. 
. (
1999
).
Antivascular endothelial growth factor receptor (fetal liver kinase 1) monoclonal antibody inhibits tumor angiogenesis and growth of several mouse and human tumors.
Cancer Res.
59
,
5209
5218
.
Raimondi
C.
,
Ruhrberg
C.
(
2013
).
Neuropilin signalling in vessels, neurons and tumours.
Semin. Cell Dev. Biol.
24
,
172
178
.
Ruhrberg
C.
,
Gerhardt
H.
,
Golding
M.
,
Watson
R.
,
Ioannidou
S.
,
Fujisawa
H.
,
Betsholtz
C.
,
Shima
D. T.
(
2002
).
Spatially restricted patterning cues provided by heparin-binding VEGF-A control blood vessel branching morphogenesis.
Genes Dev.
16
,
2684
2698
.
Smith
L. E.
,
Wesolowski
E.
,
McLellan
A.
,
Kostyk
S. K.
,
D’Amato
R.
,
Sullivan
R.
,
D’Amore
P. A.
(
1994
).
Oxygen-induced retinopathy in the mouse.
Invest. Ophthalmol. Vis. Sci.
35
,
101
111
.
Soker
S.
,
Miao
H. Q.
,
Nomi
M.
,
Takashima
S.
,
Klagsbrun
M.
(
2002
).
VEGF165 mediates formation of complexes containing VEGFR-2 and neuropilin-1 that enhance VEGF165-receptor binding.
J. Cell. Biochem.
85
,
357
368
.
Stalmans
I.
,
Ng
Y. S.
,
Rohan
R.
,
Fruttiger
M.
,
Bouché
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
.
Tomanek
R. J.
(
1996
).
Formation of the coronary vasculature: a brief review.
Cardiovasc. Res.
31
,
E46
E51
.
Valdembri
D.
,
Caswell
P. T.
,
Anderson
K. I.
,
Schwarz
J. P.
,
König
I.
,
Astanina
E.
,
Caccavari
F.
,
Norman
J. C.
,
Humphries
M. J.
,
Bussolino
F.
, et al. 
. (
2009
).
Neuropilin-1/GIPC1 signaling regulates alpha5beta1 integrin traffic and function in endothelial cells.
PLoS Biol.
7
,
e25
.
Vander Kooi
C. W.
,
Jusino
M. A.
,
Perman
B.
,
Neau
D. B.
,
Bellamy
H. D.
,
Leahy
D. J.
(
2007
).
Structural basis for ligand and heparin binding to neuropilin B domains.
Proc. Natl. Acad. Sci. USA
104
,
6152
6157
.
Vieira
J. M.
,
Schwarz
Q.
,
Ruhrberg
C.
(
2007
).
Selective requirements for NRP1 ligands during neurovascular patterning.
Development
134
,
1833
1843
.
von Wronski
M. A.
,
Raju
N.
,
Pillai
R.
,
Bogdan
N. J.
,
Marinelli
E. R.
,
Nanjappan
P.
,
Ramalingam
K.
,
Arunachalam
T.
,
Eaton
S.
,
Linder
K. E.
, et al. 
. (
2006
).
Tuftsin binds neuropilin-1 through a sequence similar to that encoded by exon 8 of vascular endothelial growth factor.
J. Biol. Chem.
281
,
5702
5710
.
Wang
L.
,
Zeng
H.
,
Wang
P.
,
Soker
S.
,
Mukhopadhyay
D.
(
2003
).
Neuropilin-1-mediated vascular permeability factor/vascular endothelial growth factor-dependent endothelial cell migration.
J. Biol. Chem.
278
,
48848
48860
.
Yaqoob
U.
,
Cao
S.
,
Shergill
U.
,
Jagavelu
K.
,
Geng
Z.
,
Yin
M.
,
de Assuncao
T. M.
,
Cao
Y.
,
Szabolcs
A.
,
Thorgeirsson
S.
, et al. 
. (
2012
).
Neuropilin-1 stimulates tumor growth by increasing fibronectin fibril assembly in the tumor microenvironment.
Cancer Res.
72
,
4047
4059
.

Competing interests

This study was financially assisted in part by Ark Therapeutics Ltd, which has had an interest in developing therapies to inhibit NRP1.

This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/3.0), which permits unrestricted use, distribution and reproduction in any medium provided that the original work is properly attributed.

Supplementary information