Skip to main content
Erschienen in: Angiogenesis 4/2018

Open Access 21.05.2018 | Brief Communication

A xenograft model for venous malformation

verfasst von: Jillian Goines, Xian Li, Yuqi Cai, Paula Mobberley-Schuman, Megan Metcalf, Steven J. Fishman, Denise M. Adams, Adrienne M. Hammill, Elisa Boscolo

Erschienen in: Angiogenesis | Ausgabe 4/2018

Abstract

Vascular malformations are defects caused by the abnormal growth of the vasculature. Among them, venous malformation (VM) is an anomaly characterized by slow-flow vascular lesions with abnormally shaped veins, typically in sponge-like configuration. VMs can expand over years causing disfigurement, obstruction of vital structures, thrombosis, bleeding, and pain. Treatments have been very limited and primarily based on supportive care, compression garments, sclerotherapy, and/or surgical resection. Sirolimus treatment has recently shown efficacy in some patients with complicated vascular anomalies, including VMs. Activating somatic TIE2 gene mutations have been identified in up to 60% of VMs and PIK3CA mutations have been found in another 25%. Here, we report a xenograft model of VM that reflects the patients’ mutation heterogeneity. First, we established a protocol to isolate and expand in culture endothelial cells (VM–EC) from VM tissue or VM blood of nine patients. In these cells, we identified somatic mutations of TIE2, PIK3CA, or a combination of both. Both TIE2 and PIK3CA mutations induced constitutive AKT activation, while TIE2 mutations also showed high MAPK–ERK signaling. Finally, VM–EC implanted into immune-deficient mice generated lesions with ectatic blood-filled channels with scarce smooth muscle cell coverage, similar to patients’ VM. This VM xenograft model could be instrumental to test the therapeutic efficacy of Sirolimus in the presence of the different TIE2 or PIK3CA mutations or to test for efficacy of additional compounds in targeting the specific mutated protein(s), thus enabling development of personalized treatment options for VM patients.
Hinweise

Electronic supplementary material

The online version of this article (https://​doi.​org/​10.​1007/​s10456-018-9624-7) contains supplementary material, which is available to authorized users.
Abkürzungen
VM
Venous malformation
EC
Endothelial cells
PI3K
Phosphoinositide 3-kinase

Introduction

Venous malformation (VM) is a congenital chronic condition that can be severely disfiguring [1, 2]. The incidence of VM is approximately 1 in 5–10,000 [2, 3]. VMs consist of endothelial-lined dilated slow-flow dysmorphic venous channels, typically in a sponge-like configuration with impaired smooth muscle cell coverage. VMs can cause significant pain, obstruction of organ function, and in some cases, localized intravascular coagulopathy resulting in risk of bleeding and thrombosis. Current treatment of VM is based on ablation of the abnormal vessels by sclerotherapy or removal by surgical excision; however, these approaches are not curative and regrowth is common [2]. In recent clinical trials, treatment with the mTOR inhibitor Sirolimus has shown efficacy in some patients affected by VMs [4, 5].
Activating germline or somatic TEK (TIE2) mutations are associated with VM [6, 7]. TIE2 is an endothelium-specific receptor tyrosine kinase that regulates both maintenance of vascular quiescence and promotion of angiogenesis. More recently, somatic mutations in the catalytic subunit of class I phosphoinositide 3-kinases (PIK3CA), reported in several types of cancer [8], overgrowth syndromes [911], and lymphatic malformations [1214], have been identified in about 20–25% of VM cases [1517].
Murine models for VM have been reported. The first model is based on subcutaneous injection of human umbilical vein endothelial cells (HUVEC) engineered to express the most frequent mutation identified in VM patients TIE2 p.L914F [5], or other TIE2 mutations [18]. The second system relies on the transgenic expression of PIK3CA p.H1047R in Sprr2f+ cells (epithelial and endothelial), in the embryonic mesoderm or in VE-Cadherin+ cells [15, 16, 19].
Here, we isolated and characterized EC from tissue or lesional blood from VM patients (VM–EC) and determined the presence of TIE2 (p.L914F), PIK3CA (p.H1047R, C420R), or combination of both (TIE2 p.R915C and PIK3CA p.Q546K) somatic mutations. We determined that the TIE2 mutation was not present in the non-endothelial cells obtained from VM samples. Furthermore, we established a xenograft model of VM by subcutaneous injection of the VM–EC. The mutated VM–EC formed enlarged blood-filled vessels with scarce smooth muscle cell coverage, akin to human VM. This model is reflective of the range of mutations found in patients.

Results

Isolation and characterization of endothelial cells from tissue and blood derived from VM lesions

EC were successfully isolated from 9 VM patients (3 solid tissues and 6 lesion blood samples collected during sclerotherapy procedure) (Table 1). VM–EC monolayers presented with a homogeneous cobblestone appearance up to passage 7 (Supplemental Fig.S1) and expressed EC-specific markers CD31, vonWillebrand Factor (vWF), and vascular endothelial (VE)-Cadherin (Fig. 1a), similarly to normal, control EC (Fig. 1b). VM–EC did not show expression of lymphatic marker Prox1 nor smooth muscle alpha actin (αSMA) (Fig. 1a). Quantitative real-time polymerase chain reaction (qRT-PCR) revealed that each VM–EC population expressed EC-specific genes at similar levels (p > 0.12) compared to human umbilical vein endothelial cells (HUVEC) and cord blood endothelial colony forming cells (cbECFC) (Fig. 1c).
Table 1
Patients’ VM samples
Patient ID
VM location
Age at collection (years)
Sex
Endothelial cell source
Date obtained
PIK3CA mutation
TIE2 mutation
Presence in blood or saliva
Mutant allelic frequency in VM–EC
1
Leg
15
F
Tissue
1/7/14
 
L914F
n/a
128/273
(46.9%)
A
Thigh
4
F
Lesion Blood
5/8/15
 
L914F
n/a
n/a
C
Hip, pelvis, perineum, leg
15
F
Lesion Blood
8/25/15
 
L914F
No
n/a
D
Lower extremity IM
18
F
Lesion Blood
10/13/15
H1047R
 
No
n/a
E
Calf IM
18
F
Lesion Blood
11/10/15
C420R
 
No
34/57
(59.6%)
9
Buttock, knee, leg
9
M
Tissue
8/22/14
 
L914F
n/a
n/a
11
Arm
3
M
Tissue
12/22/16
 
L914F
n/a
n/a
G
Chest
20
F
Lesion Blood
4/13/16
 
L914F
No
n/a
K
Thigh, calf, ankle
12
F
Lesion Blood
3/4/17
Q546K
R915C
No
n/a
IM intramuscular, n/a not available; no mutation not detected

TIE2 and PIK3CA somatic mutations exist in VM–EC

VM–EC DNA Sanger sequencing analysis was performed for TIE2 exon 17 (tyrosine kinase domain). If initial analysis did not detect a p.L914F mutation, we carried out TIE2 next-generation sequencing (NGS) to screen for other TIE2 mutations. Next, primers amplifying PIK3CA exons 7, 9 (α-helical domain), and 20 (tyrosine kinase domain) were used to further determine, by DNA Sanger sequencing, the presence of mutations frequently associated with vascular anomalies (at sites p.C420, E542, E545, and H1047) [9, 11, 1517]. TIE2 p.L914F mutations were identified in 6/9 VM–EC, making this the most frequent mutation in our study and in agreement with previous literature [6, 20]. Mutually exclusive PIK3CA mutations were present in 2/9 VM–EC (p.H1047R and C420R). Interestingly, sequencing analysis revealed a simultaneous expression of TIE2 p.R915C and PIK3CA p.Q546K mutations in 1/9 VM–EC (VMK EC) (Table 1; Fig. 2a). 7/7 single cell-derived clonal populations expressed both the TIE2 and the PIK3CA mutations, suggesting VMK EC is a pure population of double-mutant EC (Supplemental Fig. S2). Mutated allele frequencies in VM–EC that were subjected to TIE2 gene or whole exome NGS equaled 46.9% in VM1 EC (TIE2 c.2740C > T) and 59.6% in VME EC (PIK3CA c.1258T > C) (Table 1), suggesting the VM–EC populations tested contain predominantly heterozygous mutant cells. The TIE2 or PIK3CA mutation(s) were also analyzed in five available CD31 negative (CD31, non-endothelial) cell populations obtained from the same patient sample. No TIE2 p.L914F variant was detected in these CD31 cell populations (Table 2; Fig. 2b). Mutations found in VM–EC were further confirmed to be somatic by Sanger sequencing of peripheral blood or saliva genomic DNA from the same patient (Table 1).
Table 2
Mutation analysis in CD31+ and CD31 cells from VM patients
Patient ID
TIE2/ PIK3CA mutation detected
Endothelial cells (CD31+)
Non-endothelial cells (CD31)
1
TIE2 L914F
No
A
TIE2 L914F
No
9
TIE2 L914F
No
11
TIE2 L914F
No
G
TIE2 L914F
No
No: Not detected

Patient-derived VM–EC xenograft model recapitulates formation of enlarged blood vessels as seen in patients’ VM lesions

VM–EC were expanded in culture and injected subcutaneously into immune-deficient mice to determine if they could recapitulate the histological features seen in VM patient tissue. Explanted lesions were visibly vascularized at day 9 (Fig. 3a). Hematoxylin and eosin (H&E) staining of lesion sections showed numerous enlarged, blood-filled vessels similar to patient VM histology (Supplemental Fig. S3) and a VM murine model based on HUVEC–TIE2–L914F [5] (Fig. 3b). Normal HUVEC and HUVEC–TIE2–WT (wild type) injected explants did not show evidence of vascularization (Fig. 3b and Supplemental Fig. S4). VM–EC and HUVEC–TIE2–L914F explant sections showed positive staining for human-specific EC-marker Ulex europaeus agglutinin I (UEA), revealing that the lining of vascular channels consisted predominantly of the injected human-EC rather than invading mouse-EC (Fig. 3a). Isolectin B4 (IB4) staining further confirmed that enlarged blood vessels in the VM–EC explants were not mouse EC-derived (Supplemental Fig. S5, and positive/negative controls in Supplemental Fig. S6). Staining with αSMA showed scarce smooth muscle cell coverage (Fig. 3a), as previously reported in VM patient tissue [7]. Quantitative analysis of VM–EC explants at day 9 revealed a significant increase in vascular area (p < 0.01) and vascular density (p < 0.001) when compared to normal HUVEC explants (Fig. 3c). Furthermore, PIK3CA mutated VM–EC generated VM lesions with a significantly higher (p < 0.001) vascular density (Fig. 3c).

Constitutive AKT activation downstream of the TIE2 and PIK3CA mutations in VM–EC

To analyze the activation status of the TIE2–PI3KCA downstream pathways in the VM–EC with different types of mutation, we performed immunoblot analysis of phospho-TIE2, phospho-AKT, and phospho-ERK levels. VM–EC (G,9,11,K) expressing TIE2 mutations (p.L914F and p.R915C) showed constitutive activation of the TIE2 receptor that was not present in the PIK3CA mutated VM–EC (D, E) and in HUVEC (Fig. 4a). AKT activation was elevated downstream of both TIE2 (p.L914F) and PIK3CA mutations (p.H1047R, p.C420R) and was higher in VM–EC expressing PIK3CA variants. Conversely, ERK activation was higher in the TIE2 p.L914F VM–EC when compared to VM–EC expressing PIK3CA mutations (Fig. 4b). VM–EC and HUVEC responded to ANGPT1 stimulation by activating both the PI3K–AKT and the MAPK–ERK pathways. VMK EC expressing both TIE2 and PIK3CA mutations (double TIE2 p.R915C/PIK3CA p.Q546K) showed higher phospho-AKT levels when compared to both TIE2 and PIK3CA single mutation VM–EC and phospho-ERK levels were similar to the TIE2 p.L914F.

Discussion

Here, we show successful isolation and propagation of EC from both lesional blood and solid tissue types from VM patient samples. We have also determined that TIE2 mutations are specifically expressed in the EC population of the VM tissue. We generated a novel xenograft model by injecting mutated EC derived from VM patients (Fig. 5). This murine model recapitulates the patients’ VM histology and can be a very useful tool to further study the pathology of VM on a patient-to-patient basis.
The feasibility of EC isolation from VM was first demonstrated with the use of solid VM lesion tissue, including intramuscular VM samples [21, 22]. In our study, in addition to solid tissue samples, we established a protocol to isolate EC from the blood aspirated from VM lesions during sclerotherapy procedure. This fluid is normally discarded and is more frequently available than solid tissue resected by surgery, as sclerotherapy is a less invasive procedure and commonly used to treat VM patients [2, 23]. Furthermore, we have been able to detect somatic TIE2 and/or PIK3CA mutations in the EC derived from VM blood, suggesting this as a potential strategy to identify somatic mutations in patients that do not undergo surgery or solid tissue biopsy.
Somatic TIE2 mutations are detected in about 34–61.5% of VM patients and they are mostly located on the tyrosine kinase domain of the TIE2 receptor [1517, 20]. In our study, we predominantly identified the TIE2 variant p.L914F (6/9), which is reported to be the most frequent among the TIE2 mutation types [6, 20].
Activating PIK3CA mutations are associated with about 20–25% of sporadic VM cases and are reported as mutually exclusive with TIE2 mutations [1517]. Here, we show that 1/9 VM–EC expressed both a TIE2 p.R915C and a PIK3CA p.Q546K mutation and that both mutations are present in each cell within our VMK EC population. PIK3CA mutation at p.Q546 site has been previously reported as single mutation and the TIE2 mutation p.R915C as mosaic single or double TIE2 mutation in combination with p.Y897H/C [17, 18]. While these studies investigated the presence of the somatic TIE2 or PIK3CA mutation(s) in VM tissue, here we performed in-depth studies to show that the TIE2 mutation p.L914F is exclusively expressed by the EC population within the lesion. We speculate that this is true for every TIE2 and PIK3CA mutation, although we cannot confirm as we were not able to grow CD31 negative cells from all specimens obtained.
We have established that both TIE2 and PIK3CA mutations result in constitutive AKT activation in VM–EC, similarly to HUVEC overexpressing TIE2 p.L914F or PIK3CA p.H1047R [17]. Furthermore, our results determined that PIK3CA mutations are stronger AKT activators while TIE2 mutations also affect the MAPK–ERK signaling.
We have previously reported a TIE2 mutation-dependent murine model of VM based on subcutaneous injection of HUVEC–TIE2–L914F. These mice form VM lesions where the injected cells remodel into massively enlarged blood vessels that expand over time [5]. More recently, some studies reported animal models of VM consisting of transgenic mice that express PIK3CA p.H1047R in the vasculature [15, 16, 19]. These models have been instrumental to demonstrate the efficacy of mTOR inhibition in preventing VM lesion growth, but fail to reflect the heterogeneity of the mutations identified within VMs and the patients’ genetic background. In vitro, rapamycin treatment was shown to inhibit AKT activation in HUVEC expressing TIE2 p.L914F, p.R1099X or PIK3CA p.E542K, p.E545K, and p.H1047R [5, 17]. Additional in vivo studies are needed to confirm drug efficacy on the different mutation types. Here, we sought to use patient-derived VM–EC to generate a xenograft model that would be more representative of the VM population and that could be used as a platform for future testing of drug efficacy on VM lesions with a range of TIE2 and/or PIK3CA variants. All of the VM–EC we analyzed formed aberrant vascular networks when injected in vivo. In addition, we identified an association between the high levels of AKT activation in PIK3CA mutated VM–EC and vascular density in the xenograft. Furthermore, we strongly encourage future studies with a larger sample size to investigate association between number and/or diameter of the VM blood vessels and the mutation type.

Materials and methods

Tissue samples

Patient tissue samples were obtained from participants after informed consent from the Collection and Repository of Tissue Samples and Data from Patients with Tumors and Vascular Anomalies (IRB #2008-2001 per institutional policies) at Cincinnati Children’s Hospital Medical Center (CCHMC), Cancer and Blood Disease Institute (CBDI), and with approval of the Committee on Clinical Investigation at Boston Children’s Hospital. Samples include excised tissue, blood/fluid aspirated at the time of sclerotherapy, peripheral blood, and buccal mucosal swabs. Collected data and identifying names were stored in a secure database maintained by the CBDI. A unique patient number, created by the database, was assigned to each sample. This was further de-identified by creating a patient ID for use in this study.

Cell culture

Solid tissue samples were minced and digested in 5 ml of Dulbecco’s modified Eagle medium (DMEM) (Gibco) supplemented with 2% fetal bovine serum (FBS) (Hyclone), 1X Ca2+/Mg2+, and 1 mg/ml of collagenase A (Roche) dissolved in phosphate-buffered saline solution (PBS) (Thermofisher) at 37 °C for 30 min. The digested tissue was homogenized with a pestle with 5 ml of PBS supplemented with 0.5% bovine serum albumin (BSA) (Sigma) and 2% penicillin–streptomycin–glutamine (PSG) (Corning) four times. Homogenized tissue was then filtered through a 100 µm cell strainer to remove fragments, followed by centrifugation for 5 min at 1500 rpm. Lesional blood was diluted in PBS to a final volume of 40 ml and centrifuged at 1000 rpm for 5 min. Tissue culture plates, 100 mm, were coated with 1 µg/cm2 of fibronectin (Millipore) in 0.1 M Na2CO3, pH 9.4, coating buffer, incubated at 37 °C, 5% CO2 for 20 min, and washed with PBS. The pellet from both tissue or fluid samples was re-suspended in Endothelial Growth Medium-2 (EGM-2) (Lonza), with 20% FBS, 1% PSG, and seeded onto fibronectin-coated plates. Colonies of endothelial cells appeared evident within 2–3 weeks from cell seeding. When cells reached 80% confluency, they were trypsinized and pelleted by centrifugation for 5 min at 1500 rpm and EC were purified with anti-CD31 antibody-conjugated magnetic beads (Dynal) following the manufacturer’s instructions. Retrovirally transfected HUVEC–TIE2–WT or TIE2–L914F were obtained as previously described [5, 24], in brief: full-length TIE2–WT or TIE2–L914F were cloned into pMXs vector and packaging cell line 293-GPG VSV-G was transfected for retrovirus production with Fugene 6 (Roche). All cell types were cultured in identical conditions: EGM-2 20% FBS, 1% PSG, on fibronectin-coated plates at 37 °C, 5% CO2.

Immunocytochemistry

Photos of cell monolayers, passage 3–7, were taken with a phase-contrast microscope (Zeiss) using ZenLite Software. Immunocytochemistry was performed when cells reached 80–90% confluency. Cells were fixed with cold methanol (Fisher Chemicals) at 4 °C for 10 min and blocked in 5% horse serum (Vector Laboratories) in PBS. Primary antibody incubation with anti-CD31 (1:50, Dako), vonWillebrand Factor (vWF) (1:100, Dako), αSMA (1:500, Sigma), VE-Cadherin (1:50, Santa Cruz), and PROX1 (1:50, R&D Systems) was performed for 1 h at room temperature (RT). Fluorescein isothiocyanate (FITC)-conjugated secondary antibodies (1:200, Vector Laboratories) were used for 1 h at RT. Samples were mounted using Prolong Gold with 4′,6-diamidino-2-phenylindole (DAPI) (Life Technologies) and images acquired using C2 confocal microscope (Nikon).

cDNA synthesis and quantitative RT-PCR

VM–EC, cbECFC, and HUVEC monolayers at passage 3 or 4 were lysed and homogenized using Qiashredder (Qiagen) and RNA isolated with RNeasy Minikit (Qiagen). RNA concentration and quality were determined with Nanodrop 2000c Spectrophotometer (Thermofisher). Synthesis of cDNA was performed using 1.0 µg of RNA and iScript cDNA Synthesis Kit (Biorad). Quantitative real-time PCR (qRT-PCR) was performed using 10 ng of cDNA in 20 µl reaction using SsoAdvanced Universal SYBR Green Supermix (Biorad) and a CFX96 Real-Time PCR Detection System (Biorad). Thermocycler parameters were 95 °C for 30 s followed by 40 cycles of 95 °C for 10 s, 60 °C for 30 s with plate reading, and a 65–95 °C melt curve in 5 s increments. Gene expression was calculated by standard curve method using target gene expression relative to 18S mRNA. PCR primer sequences are listed in Supplemental Table S1.

DNA sequencing

DNA was extracted from VM–EC (CD31+) and non-endothelial cell population (CD31) using QIAamp DNA Mini Kit (Qiagen). Genomic DNA from peripheral blood, plasma, serum, or buccal swab was extracted using Purelink Genomic DNA Mini Kit (Invitrogen). DNA quality and quantity were determined using a Nanodrop 2000c Spectrophotometer. 200 ng of DNA was used for PCR using GoTaq Polymerase Master Mix (Promega). TIE2 (TEK) coding sequence next-generation sequencing (NGS) and analysis was performed at the CCHMC DNA Sequencing and Genotyping Core following GeneRead Panel (Qiagen) sequence amplification.
To confirm TIE2 mutations and identify PIK3CA mutations, DNA Sanger sequencing was performed as follows. Primers were used to amplify TIE2 exon 17 and PIK3CA exons 7, 9, and 20 (Integrated DNA Technologies) (Supplemental Table S1). Amplification product was purified using QIAquick Gel Extraction Kit (Qiagen) and sequenced at CCHMC DNA Sequencing and Genotyping Core. Electropherogram peak results were visualized using CodonCode Aligner (CodonCode Corporation).

Single cell-derived clone mutation analysis

VMK EC clones were obtained by diluting 1 cell/400 µl of EGM-2 20%FBS, 1% PSG medium, and 100 µL were seeded into each fibronectin-coated well of 96-well plates. Single colonies were then isolated and propagated until efficient cell numbers were met for DNA extraction, PCR amplification, and Sanger sequencing for mutation analysis.

Xenograft model for VM

After cell expansion, 2.5–3.5 × 106 VM–EC, HUVEC, HUVEC–TIE2–L914F, and HUVEC–TIE2–WT were suspended in Matrigel™ (Corning) and injected subcutaneously on both flanks of 6- to 7-week-old male athymic nu/nu mice (Envigo). Lesions were dissected after 9 days, fixed in 10% formalin, and paraffin embedded. After Hematoxylin and Eosin (H&E) staining, five images were taken randomly per section using EVOS (Life Technologies), followed by vascular density (vessels/mm2) and vascular area (%) quantification with ImageJ software.

Immunostaining

Paraffin tissue sections were stained following antigen retrieval with citrate buffer, pH 6.0, and blocking using 5% horse serum in PBS. Immunohistochemistry was performed with biotin-conjugated Ulex europaeus I (UEA)(1:100, Vector Laboratories). Peroxidase was quenched using 3% H2O2(Sigma) prior incubation with streptavidin-conjugated horseradish peroxidase (HRP) (1:200, Vector Laboratories) followed by diaminobenzidine (DAB) (Vector Laboratories). Samples were mounted with VectaMount (Vector Laboratories). Immunofluorescence was performed using UEA, Griffonia simplicifolia Isolectin B4 (IB4) (1:50 and 1:100, Vector Laboratories), and anti-αSMA (1:500, Sigma). FITC/ Texas Red-conjugated secondary antibodies (1:200, Vector Laboratories) were used. Samples were mounted using Prolong Gold with DAPI. Images were acquired using Nikon C2 confocal microscope and NIS-Elements C imaging (Nikon).

Immunoblotting

VM–EC and HUVEC were starved for 3 h in Endothelial Basal Medium-2 (EBM-2) supplemented with 1% FBS and then treated with Angiopoietin 1 (ANGPT1) 1 µg/ml (R&D Systems) for 15 min. Cells were then washed with PBS then lysed using RIPA buffer (Boston Bioproduct) with phosphatase inhibitor cocktail (Roche). Lysates were subjected to SDS-PAGE and transferred to Immobilon-P membrane. Membranes were probed with antibodies against the following: phospho-TIE2(Y992), TIE2, phospho-AKT (Ser473), AKT, phospho-ERK 1/2, ERK 1/2 (all 1:1000 Cell Signaling Technologies), and β-actin (1:2000, Sigma Aldrich). Membranes were incubated with peroxidase-conjugated secondary antibodies (1:5000, Vector Laboratories). Antigen–antibody complexes were visualized using ECL (Pierce) and chemiluminescent sensitive film. Band intensity was analyzed with ImageJ software.

Statistical analysis

Results are reported as means ± standard deviation (SD). Unpaired student t test was performed, and p values reported with differences between groups considered significant at p ≤ 0.05.

Acknowledgements

Research reported in this manuscript was supported by the National Heart, Lung, and Blood Institute, under Award Number R01 HL117952 (E.B.), part of the National Institutes of Health. This project was also supported by the Charles H. Hood Foundation (E.B.) and by the National Center for Advancing Translational Sciences of the National Institutes of Health, under Award Number 5UL1TR001425-03 (E.B.). The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health. We thank Drs. Peter Dickie for providing VM9 EC, Joyce Bischoff for assistance with the collection of the VM1 tissue sample and for providing cbECFC, Philip Dexheimer for the TIE2 NGS analysis, Suhas Kallapur and Pietro Presicce for assisting with HUVEC isolation, and the Pathology Research Core (PATH) and DNA Sequencing Core (DNA) at CCHMC.

Compliance with ethical standards

Conflict of interest

The authors declare that they have no conflicts of interest.

Ethical approval

All procedures performed in studies involving human participants were in accordance with the ethical standards of the institutional and/or national research committee and with the 1964 Helsinki declaration and its later amendments or comparable ethical standards. All applicable international, national, and/or institutional guidelines for the care and use of animals were followed. All procedures performed in studies involving animals were in accordance with the ethical standards of the institution or practice at which the studies were conducted.
Open AccessThis article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://​creativecommons.​org/​licenses/​by/​4.​0/​), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

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.

Anhänge

Electronic supplementary material

Below is the link to the electronic supplementary material.
Literatur
1.
Zurück zum Zitat Burrows PE, Mason KP (2004) Percutaneous treatment of low flow vascular malformations. J Vasc Interv Radiol 15(5):431–445CrossRef Burrows PE, Mason KP (2004) Percutaneous treatment of low flow vascular malformations. J Vasc Interv Radiol 15(5):431–445CrossRef
4.
Zurück zum Zitat Adams DM, Trenor CC 3rd, Hammill AM, Vinks AA, Patel MN, Chaudry G, Wentzel MS, Mobberley-Schuman PS, Campbell LM, Brookbank C, Gupta A, Chute C, Eile J, McKenna J, Merrow AC, Fei L, Hornung L, Seid M, Dasgupta AR, Dickie BH, Elluru RG, Lucky AW, Weiss B, Azizkhan RG (2016) Efficacy and Safety of sirolimus in the treatment of complicated vascular anomalies. Pediatrics 137(2):e20153257. https://doi.org/10.1542/peds.2015-3257 CrossRefPubMedPubMedCentral Adams DM, Trenor CC 3rd, Hammill AM, Vinks AA, Patel MN, Chaudry G, Wentzel MS, Mobberley-Schuman PS, Campbell LM, Brookbank C, Gupta A, Chute C, Eile J, McKenna J, Merrow AC, Fei L, Hornung L, Seid M, Dasgupta AR, Dickie BH, Elluru RG, Lucky AW, Weiss B, Azizkhan RG (2016) Efficacy and Safety of sirolimus in the treatment of complicated vascular anomalies. Pediatrics 137(2):e20153257. https://​doi.​org/​10.​1542/​peds.​2015-3257 CrossRefPubMedPubMedCentral
5.
Zurück zum Zitat Boscolo E, Limaye N, Huang L, Kang KT, Soblet J, Uebelhoer M, Mendola A, Natynki M, Seront E, Dupont S, Hammer J, Legrand C, Brugnara C, Eklund L, Vikkula M, Bischoff J, Boon LM (2015) Rapamycin improves TIE2-mutated venous malformation in murine model and human subjects. J Clin Invest 125(9):3491–3504. https://doi.org/10.1172/JCI76004 CrossRefPubMedPubMedCentral Boscolo E, Limaye N, Huang L, Kang KT, Soblet J, Uebelhoer M, Mendola A, Natynki M, Seront E, Dupont S, Hammer J, Legrand C, Brugnara C, Eklund L, Vikkula M, Bischoff J, Boon LM (2015) Rapamycin improves TIE2-mutated venous malformation in murine model and human subjects. J Clin Invest 125(9):3491–3504. https://​doi.​org/​10.​1172/​JCI76004 CrossRefPubMedPubMedCentral
6.
Zurück zum Zitat Limaye N, Wouters V, Uebelhoer M, Tuominen M, Wirkkala R, Mulliken JB, Eklund L, Boon LM, Vikkula M (2009) Somatic mutations in angiopoietin receptor gene TEK cause solitary and multiple sporadic venous malformations. Nat Genet 41(1):118–124CrossRef Limaye N, Wouters V, Uebelhoer M, Tuominen M, Wirkkala R, Mulliken JB, Eklund L, Boon LM, Vikkula M (2009) Somatic mutations in angiopoietin receptor gene TEK cause solitary and multiple sporadic venous malformations. Nat Genet 41(1):118–124CrossRef
7.
Zurück zum Zitat Vikkula M, Boon LM, Carraway KL 3rd, Calvert JT, Diamonti AJ, Goumnerov B, Pasyk KA, Marchuk DA, Warman ML, Cantley LC, Mulliken JB, Olsen BR (1996) Vascular dysmorphogenesis caused by an activating mutation in the receptor tyrosine kinase TIE2. Cell 87(7):1181–1190CrossRef Vikkula M, Boon LM, Carraway KL 3rd, Calvert JT, Diamonti AJ, Goumnerov B, Pasyk KA, Marchuk DA, Warman ML, Cantley LC, Mulliken JB, Olsen BR (1996) Vascular dysmorphogenesis caused by an activating mutation in the receptor tyrosine kinase TIE2. Cell 87(7):1181–1190CrossRef
8.
Zurück zum Zitat Samuels Y, Wang Z, Bardelli A, Silliman N, Ptak J, Szabo S, Yan H, Gazdar A, Powell SM, Riggins GJ, Willson JK, Markowitz S, Kinzler KW, Vogelstein B, Velculescu VE (2004) High frequency of mutations of the PIK3CA gene in human cancers. Science 304(5670):554. https://doi.org/10.1126/science.1096502 CrossRef Samuels Y, Wang Z, Bardelli A, Silliman N, Ptak J, Szabo S, Yan H, Gazdar A, Powell SM, Riggins GJ, Willson JK, Markowitz S, Kinzler KW, Vogelstein B, Velculescu VE (2004) High frequency of mutations of the PIK3CA gene in human cancers. Science 304(5670):554. https://​doi.​org/​10.​1126/​science.​1096502 CrossRef
11.
Zurück zum Zitat Riviere JB, Mirzaa GM, O’Roak BJ, Beddaoui M, Alcantara D, Conway RL, St-Onge J, Schwartzentruber JA, Gripp KW, Nikkel SM, Worthylake T, Sullivan CT, Ward TR, Butler HE, Kramer NA, Albrecht B, Armour CM, Armstrong L, Caluseriu O, Cytrynbaum C, Drolet BA, Innes AM, Lauzon JL, Lin AE, Mancini GM, Meschino WS, Reggin JD, Saggar AK, Lerman-Sagie T, Uyanik G, Weksberg R, Zirn B, Beaulieu CL, Finding of Rare Disease Genes Canada C, Majewski J, Bulman DE, O’Driscoll M, Shendure J, Graham JM Jr, Boycott KM, Dobyns WB (2012) De novo germline and postzygotic mutations in AKT3, PIK3R2 and PIK3CA cause a spectrum of related megalencephaly syndromes. Nat Genet 44(8):934–940. https://doi.org/10.1038/ng.2331 CrossRefPubMedPubMedCentral Riviere JB, Mirzaa GM, O’Roak BJ, Beddaoui M, Alcantara D, Conway RL, St-Onge J, Schwartzentruber JA, Gripp KW, Nikkel SM, Worthylake T, Sullivan CT, Ward TR, Butler HE, Kramer NA, Albrecht B, Armour CM, Armstrong L, Caluseriu O, Cytrynbaum C, Drolet BA, Innes AM, Lauzon JL, Lin AE, Mancini GM, Meschino WS, Reggin JD, Saggar AK, Lerman-Sagie T, Uyanik G, Weksberg R, Zirn B, Beaulieu CL, Finding of Rare Disease Genes Canada C, Majewski J, Bulman DE, O’Driscoll M, Shendure J, Graham JM Jr, Boycott KM, Dobyns WB (2012) De novo germline and postzygotic mutations in AKT3, PIK3R2 and PIK3CA cause a spectrum of related megalencephaly syndromes. Nat Genet 44(8):934–940. https://​doi.​org/​10.​1038/​ng.​2331 CrossRefPubMedPubMedCentral
13.
Zurück zum Zitat Luks VL, Kamitaki N, Vivero MP, Uller W, Rab R, Bovee JV, Rialon KL, Guevara CJ, Alomari AI, Greene AK, Fishman SJ, Kozakewich HP, Maclellan RA, Mulliken JB, Rahbar R, Spencer SA, Trenor CC 3rd, Upton J, Zurakowski D, Perkins JA, Kirsh A, Bennett JT, Dobyns WB, Kurek KC, Warman ML, McCarroll SA, Murillo R (2015) Lymphatic and other vascular malformative/overgrowth disorders are caused by somatic mutations in PIK3CA. J Pediatr 166(4):1048–1054.e1041–1045. https://doi.org/10.1016/j.jpeds.2014.12.069 CrossRefPubMedPubMedCentral Luks VL, Kamitaki N, Vivero MP, Uller W, Rab R, Bovee JV, Rialon KL, Guevara CJ, Alomari AI, Greene AK, Fishman SJ, Kozakewich HP, Maclellan RA, Mulliken JB, Rahbar R, Spencer SA, Trenor CC 3rd, Upton J, Zurakowski D, Perkins JA, Kirsh A, Bennett JT, Dobyns WB, Kurek KC, Warman ML, McCarroll SA, Murillo R (2015) Lymphatic and other vascular malformative/overgrowth disorders are caused by somatic mutations in PIK3CA. J Pediatr 166(4):1048–1054.e1041–1045. https://​doi.​org/​10.​1016/​j.​jpeds.​2014.​12.​069 CrossRefPubMedPubMedCentral
16.
Zurück zum Zitat Castillo SD, Tzouanacou E, Zaw-Thin M, Berenjeno IM, Parker VE, Chivite I, Mila-Guasch M, Pearce W, Solomon I, Angulo-Urarte A, Figueiredo AM, Dewhurst RE, Knox RG, Clark GR, Scudamore CL, Badar A, Kalber TL, Foster J, Stuckey DJ, David AL, Phillips WA, Lythgoe MF, Wilson V, Semple RK, Sebire NJ, Kinsler VA, Graupera M, Vanhaesebroeck B (2016) Somatic activating mutations in Pik3ca cause sporadic venous malformations in mice and humans. Sci Transl Med 8(332):332ra343. https://doi.org/10.1126/scitranslmed.aad9982 CrossRef Castillo SD, Tzouanacou E, Zaw-Thin M, Berenjeno IM, Parker VE, Chivite I, Mila-Guasch M, Pearce W, Solomon I, Angulo-Urarte A, Figueiredo AM, Dewhurst RE, Knox RG, Clark GR, Scudamore CL, Badar A, Kalber TL, Foster J, Stuckey DJ, David AL, Phillips WA, Lythgoe MF, Wilson V, Semple RK, Sebire NJ, Kinsler VA, Graupera M, Vanhaesebroeck B (2016) Somatic activating mutations in Pik3ca cause sporadic venous malformations in mice and humans. Sci Transl Med 8(332):332ra343. https://​doi.​org/​10.​1126/​scitranslmed.​aad9982 CrossRef
20.
Zurück zum Zitat Soblet JLN, Uebelhoer M, Boon LM, Vikkula M (2013) Variable somatic TIE2 mutations in half of sporadic venous malformations. Mol Syndromol 4(4):179–183PubMedPubMedCentral Soblet JLN, Uebelhoer M, Boon LM, Vikkula M (2013) Variable somatic TIE2 mutations in half of sporadic venous malformations. Mol Syndromol 4(4):179–183PubMedPubMedCentral
24.
Zurück zum Zitat Ory DS, Neugeboren BA, Mulligan RC (1996) A stable human-derived packaging cell line for production of high titer retrovirus/vesicular stomatitis virus G pseudotypes. Proc Natl Acad Sci USA 93(21):11400–11406CrossRef Ory DS, Neugeboren BA, Mulligan RC (1996) A stable human-derived packaging cell line for production of high titer retrovirus/vesicular stomatitis virus G pseudotypes. Proc Natl Acad Sci USA 93(21):11400–11406CrossRef
Metadaten
Titel
A xenograft model for venous malformation
verfasst von
Jillian Goines
Xian Li
Yuqi Cai
Paula Mobberley-Schuman
Megan Metcalf
Steven J. Fishman
Denise M. Adams
Adrienne M. Hammill
Elisa Boscolo
Publikationsdatum
21.05.2018
Verlag
Springer Netherlands
Erschienen in
Angiogenesis / Ausgabe 4/2018
Print ISSN: 0969-6970
Elektronische ISSN: 1573-7209
DOI
https://doi.org/10.1007/s10456-018-9624-7

Weitere Artikel der Ausgabe 4/2018

Angiogenesis 4/2018 Zur Ausgabe

Leitlinien kompakt für die Innere Medizin

Mit medbee Pocketcards sicher entscheiden.

Seit 2022 gehört die medbee GmbH zum Springer Medizin Verlag

Update Innere Medizin

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