Vascular endothelial growth factor (VEGF) and its two endothelial cell-specific receptor tyrosine kinases, Flk-1/KDR and Flt-1, play a key role in physiological and pathological angiogenesis. Hypoxia has been shown to be a major mechanism for up-regulation of VEGF and its receptors in vivo. When we exposed human umbilical vein endothelial cells to hypoxic conditions in vitro, we observed increased levels of Flt-1expression. In contrast, Flk-1/KDR mRNA levels were unchanged or slightly repressed. These findings suggest a differential transcriptional regulation of the two receptors by hypoxia. To identify regulatory elements involved in the hypoxic response, promoter regions of the mouse Flt-1 and Flk-1/KDR genes were isolated and tested in conjunction with luciferase reporter gene. In transient transfection assays, hypoxia led to strong transcriptional activation of the Flt-1 promoter, whereasFlk-1/KDR transcription was essentially unchanged. Promoter deletion analysis demonstrated a 430-bp region of the Flt-1promoter to be required for transcriptional activation in response to hypoxia. This region includes a heptamer sequence matching the hypoxia-inducible factor-1 (HIF) consensus binding site previously found in other hypoxia-inducible genes such as the VEGFgene and erythropoietin gene. We further narrowed down the element mediating the hypoxia response to a 40-base pair sequence including the putative HIF binding site. We show that this element acts like an enhancer, since it activated transcription irrespective of its location or orientation in the construct. Furthermore, mutations within the putative HIF consensus binding site lead to impaired transcriptional activation by hypoxia. These findings indicate that, unlike the KDR/Flk-1 gene, the Flt-1 receptor gene is directly up-regulated by hypoxia via a hypoxia-inducible enhancer element located at positions −976 to −937 of theFlt-1 promoter. Vascular endothelial growth factor (VEGF) and its two endothelial cell-specific receptor tyrosine kinases, Flk-1/KDR and Flt-1, play a key role in physiological and pathological angiogenesis. Hypoxia has been shown to be a major mechanism for up-regulation of VEGF and its receptors in vivo. When we exposed human umbilical vein endothelial cells to hypoxic conditions in vitro, we observed increased levels of Flt-1expression. In contrast, Flk-1/KDR mRNA levels were unchanged or slightly repressed. These findings suggest a differential transcriptional regulation of the two receptors by hypoxia. To identify regulatory elements involved in the hypoxic response, promoter regions of the mouse Flt-1 and Flk-1/KDR genes were isolated and tested in conjunction with luciferase reporter gene. In transient transfection assays, hypoxia led to strong transcriptional activation of the Flt-1 promoter, whereasFlk-1/KDR transcription was essentially unchanged. Promoter deletion analysis demonstrated a 430-bp region of the Flt-1promoter to be required for transcriptional activation in response to hypoxia. This region includes a heptamer sequence matching the hypoxia-inducible factor-1 (HIF) consensus binding site previously found in other hypoxia-inducible genes such as the VEGFgene and erythropoietin gene. We further narrowed down the element mediating the hypoxia response to a 40-base pair sequence including the putative HIF binding site. We show that this element acts like an enhancer, since it activated transcription irrespective of its location or orientation in the construct. Furthermore, mutations within the putative HIF consensus binding site lead to impaired transcriptional activation by hypoxia. These findings indicate that, unlike the KDR/Flk-1 gene, the Flt-1 receptor gene is directly up-regulated by hypoxia via a hypoxia-inducible enhancer element located at positions −976 to −937 of theFlt-1 promoter. The growth of new blood vessels (angiogenesis) is essential for embryonic development and other physiologic processes such as bone remodeling, wound healing, and ovarian cycle (1Folkman J. Klagsbrun M. Science. 1987; 235: 442-447Crossref PubMed Scopus (4046) Google Scholar, 2Folkman J. J. Natl. Cancer Inst. 1987; 82: 4-6Crossref Scopus (4409) Google Scholar). Angiogenesis is also a critical component of tumors, inflammatory arthritis, intraocular neovascular syndromes, and other disorders (3Folkman J. Hanahan D. Princess Takamatsu Symp. 1991; 22: 339-347PubMed Google Scholar, 4Kim K.J. Li B. Winer J. Armanini M. Gillett N. Phillips H.S. Ferrara N. Nature. 1993; 362: 841-844Crossref PubMed Scopus (3353) Google Scholar). The search for potential regulators of angiogenesis led to a number of candidates (aFGF, basic fibroblast growth factor, transforming growth factor-α, transforming growth factor-β, etc.) (5Folkman J. Shing Y. J. Biol. Chem. 1992; 267: 10931-10934Abstract Full Text PDF PubMed Google Scholar). Among those, VEGF 1The abbreviations used are: VEGF, vascular endothelial growth factor; HIF, hypoxia-inducible factor-1; EPO, erythropoietin; HUVE, human umbilical vein endothelial; kb, kilobase pair; bp, base pair; EGM, endothelial growth medium; EGF, epidermal growth factor; RLU, relative light units; PCR, polymerase chain reaction; RT-PCR, reverse transcriptase-polymerase chain reaction; HIE, hypoxia inducible element. and its two receptors, Flt-1 and Flk-1/KDR have been shown to be crucially involved in physiological and pathological regulation of blood vessel growth (6Ferrara N. Davis-Smtyh T. Endocr. Rev. 1997; 18: 4-25Crossref PubMed Scopus (3668) Google Scholar). Recently, it has been shown that oxygen tension plays a major role in the regulation of VEGF gene expression (7Minchenko A. Bauer T. Salceda S. Caro J. Lab. Invest. 1994; 71: 374-379PubMed Google Scholar, 8Shweiki D. Itin A. Soffer D. Keshet E. Nature. 1992; 359: 843-845Crossref PubMed Scopus (4163) Google Scholar, 9Shima D.T. Adamis A.P. Ferrara N. Yeo K.T. Yeo T.K. Allende R. Folkman J. D'Amore P.A. Mol. Med. 1995; 1: 182-193Crossref PubMed Google Scholar). VEGF mRNA expression is rapidly and reversibly induced by exposure to low oxygen conditions in a variety of normal and transformed cells. A 47-bp regulatory element located about 1 kb upstream to the VEGF transcription initiation site was found to be involved in the activation of VEGF transcription in hypoxic cells. This element includes a binding site for the transcription factor hypoxia-inducible factor-1 (HIF-1) (10Jiang B.-H. Rue E. Wang G.L. Roe R. Semenza G.L. J. Biol. Chem. 1996; 271: 17771-17778Abstract Full Text Full Text PDF PubMed Scopus (902) Google Scholar). When reporter constructs containing the VEGF sequences that mediate hypoxia inducibility were co-transfected with expression vectors encoding HIF-1 subunits, reporter gene transcription was much greater than that observed in cells transfected with the reporter alone, both in hypoxic and normoxic conditions (11Forsythe J.A. Jiang B.H. Iyer N.V. Agani F. Leung S.W. Koos R.D. Semenza G.L. Mol. Cell. Biol. 1996; 16: 4604-4613Crossref PubMed Scopus (3217) Google Scholar). HIF-1 has been shown to be involved also in the regulation of the human and mouse erythropoietin (EPO) genes (12Madan A. Curtin P.T. Proc. Natl. Acad. Sci. U. S. A. 1993; 90: 3928-3932Crossref PubMed Scopus (108) Google Scholar, 13Semenza G.L. Nejfelt M. Chi S.M. Antonarakis S.E. Proc. Natl. Acad. Sci. U. S. A. 1991; 88: 5680-5684Crossref PubMed Scopus (719) Google Scholar) as well as other hypoxia inducible genes such as the glycolytic enzymes (14Firth J.D. Ebert B.L. Pugh C.W. Ratcliffe P.J. Proc. Natl. Acad. Sci. U. S. A. 1994; 91: 6496-6500Crossref PubMed Scopus (446) Google Scholar, 15Semenza G.L. Roth P.H. Fang H.-M. Wang G.L. J. Biol. Chem. 1994; 269: 23757-23763Abstract Full Text PDF PubMed Google Scholar). Hypoxia has been proposed to play an important role also in the regulation of VEGF receptor gene expression. Exposure of rats to acute or chronic hypoxia led to pronounced up-regulation of bothFlt-1 and Flk-1/KDR genes in the lung vasculature (16Tuder R.M. Flook B.E. Voelkel N.F. J. Clin. Invest. 1995; 95: 1798-1807Crossref PubMed Scopus (532) Google Scholar). Also, Flk-1/KDR and Flt-1 mRNAs were substantially up-regulated throughout the heart following myocardial infarction in the rat (17Li J. Brown L.F. Hibberd M.G. Grossman J.D. Morgan J.P. Simons M. Am. J. Physiol. 1996; 270: H1803-H1811PubMed Google Scholar). Furthermore, Flt-1 and Flk-1/KDR mRNAs are markedly up-regulated in ischemic regions of tumors such as glioblastoma multiforme (8Shweiki D. Itin A. Soffer D. Keshet E. Nature. 1992; 359: 843-845Crossref PubMed Scopus (4163) Google Scholar, 18Plate K.H. Breier G. Weich H.A. Risau W. Nature. 1992; 359: 845-848Crossref PubMed Scopus (2122) Google Scholar, 19Phillips H.S. Armanini M. Stavrou D. Ferrara N. Westphal M. Int. J. Oncol. 1993; 2: 913-919PubMed Google Scholar). However, in vitro studies have yielded conflicting findings. Although Thieme et al. (20Thieme H. Aiello L.P. Ferrara N. King G.L. Diabetes. 1995; 44: 98-103Crossref PubMed Google Scholar) have shown that hypoxia increases VEGF receptor number by 50% in cultured bovine retinal capillary endothelial cells, the levels of Flk-1/KDR mRNA appeared to be down-regulated. Also, while an up-regulation of Flt-1 mRNA in response to hypoxia was found in cultured pericytes (21Takagi H. King G.L. Aiello L.P. Diabetes. 1996; 45: 1016-1023Crossref PubMed Scopus (116) Google Scholar) or in microvessels in skin explants (22Detmar M. Brown L.F. Berse B. Jackman R.W. Elicker R.W. Dvorak H.F. Claffey K.P. J. Invest. Dermatol. 1997; 108: 263-268Abstract Full Text PDF PubMed Scopus (236) Google Scholar), others failed to detect a similar up-regulation of Flt-1 in human umbilical vein endothelial (HUVE) cells (23Waltenberger J. Mayr U., S., P. Hombach V. Circulation. 1996; 94: 1647-1654Crossref PubMed Scopus (232) Google Scholar). Furthermore, Brogiet al. (24Brogi E. Schatteman G. Wu T. Kim E.A. Varticovski L. Keyt B. Isner J.M. J. Clin. Invest. 1996; 97: 469-476Crossref PubMed Scopus (341) Google Scholar) reported that the Flk-1/KDR mRNA is not directly induced by hypoxia in HUVE cells or in microvascular endothelial cells. It has been suggested that the in vivoup-regulation of Flk-1/KDR receptor expression is mediated by a paracrine factor released by ischemic tissues (24Brogi E. Schatteman G. Wu T. Kim E.A. Varticovski L. Keyt B. Isner J.M. J. Clin. Invest. 1996; 97: 469-476Crossref PubMed Scopus (341) Google Scholar) or by post-transcriptional mechanisms such as increased mRNA stability (23Waltenberger J. Mayr U., S., P. Hombach V. Circulation. 1996; 94: 1647-1654Crossref PubMed Scopus (232) Google Scholar). Using real time RT-PCR technology, we found an up-regulation of theFlt-1 expression compared with the Flk-1/KDR in HUVE cells. Sequence analysis of the mouse and human Flt-1 promoter revealed a heptamer element highly homologous to the HIF consensus sites present in the 5′-region of mouse and human VEGF genes and the 3′-enhancer of the human EPO gene. So far, such a homologous element could be found neither in the human nor in the mouseFlk-1 promoter sequences available. In the present study, we provide evidence that a 40-bp sequence including such an element is sufficient to confer hypoxia inducibility when tested in a heterologous promoter context and revealed enhancer-like features. These differences in the regulation further emphasize the different roles played by these receptors in mediating VEGF biological responses. A 129Sv/ev male genomic library in λGEM-11 vector (Promega, Madison, WI), generated by partial Sau3A digestion, was screened with a set of overlapping oligonucleotides covering the 5′-end of the leader sequences of the mouse Flt-1 andFlk-1 gene, respectively (25Sambrook J. Fritsch E.F. Maniatis T. Molecular Cloning: A Laboratory Manual.2nd Ed. Cold Spring Harbor Laboratory, Cold Spring Harbor, NY1989Google Scholar). Oligonucleotide probes used to isolate the mouse Flt-1 gene were mFlt-1 probe 1 and probe 2. For the mouse Flk-1 gene, oligonucleotide mFlk-1 probe 1 and probe 2 were radioactively labeled after annealing in a Klenow fill-in reaction (DNA Polymerase I, Large (Klenow) fragment (25Sambrook J. Fritsch E.F. Maniatis T. Molecular Cloning: A Laboratory Manual.2nd Ed. Cold Spring Harbor Laboratory, Cold Spring Harbor, NY1989Google Scholar). For sequences of oligonucleotides, see Table I. All enzymes used were from New England Biolabs (Beverly, MA), unless otherwise indicated. 16 pmol of each oligonucleotide was present in 20 μl of 10 mmTris-HCl, 10 mm MgCl2, 50 mm NaCl, 1 mm dithiothreitol, heated for 10 min at 55 °C, and allowed to cool for 10 min at room temperature. The mixture was made 100 μm for dGTP, dTTP, and dATP. 2 units of Klenow fragment were then added, and the total volume was adjusted to 50 μl. The labeling reaction occurred at 37 °C for 30 min in the presence of 50 μCi of [32P]dCTP (3000 mCi/mmol, Amersham Corp. Three overlapping but not identical clones of each gene were isolated. The restriction maps of these genes were generated using the partial restriction method (25Sambrook J. Fritsch E.F. Maniatis T. Molecular Cloning: A Laboratory Manual.2nd Ed. Cold Spring Harbor Laboratory, Cold Spring Harbor, NY1989Google Scholar) (data not shown). Fragments were cloned into pBluescript KS (Stratagene, La Jolla, CA). A 3.5-kb NcoI fragment of the Flt-1 gene and a 2.1-kb PstI fragment from the Flk-1 phages were subcloned into pGEM5 or pSK to generate pGEM5Flt-1(NcoI) and pSKFlk-1(PstI), respectively. Sequencing reactions were performed in an automatic sequencer (model 373A, Applied BioSystems, Foster City, CA). Both strands of the Flt-1 promoter including the region from −3181 to +352 and the region between −1824 and +148 of the Flk-1 promoter were sequenced by cycle sequencing. Sequence analysis was done using the Sequencher 3.1 program (Gene Codes Corporation, Ann Arbor, MI).Table ISequences of oligonucleotide primers and real time RT-PCR probesmFlt-1 probe 1GGT CAG CTG CTG GGA CAC CGC GGT CTT GCC TTA CGC GCT GCT CGG GTG TCT GCmFlt-1 probe 2GGC ACT TTT AAC TTC GAC CCT GAG CCA TAT CCT GTG AGA AGC AGA CAC CCG AGC AGC GCGmFlk-1 probe 1GGG GCC ATA CCG CCT CTG TGA CTT CTT TGC GGG CCA GGG ACG GAG AAG GAG TCT GmFlk-1 probe 2CTC CCT GGG CAC AGA GCC CAG TTT CTC AGG CAC AGA CTC CTT CTC CGT CCCmFlt-HIE30–1TCG CCA ATT GAG GAA CAA CGT GGA ATT AGT GTC ATGmFlt-HIE30–2TCG ACA TGA CAC TAA TTC CAC GTT GTT CCT CAA TTGmFlt-HIE40–1GAT CCT GCA TAA TTG AGG AAC AAC GTG GAA TTA GTG TCA TCG TAA GmFlt-1HIE40–2TCG ACT TAC GAT GAC ACT AAT TCC ACG TTG TTC CTC AAT TAT GCA GmFlt-HIE50–1TCG AGA TGG ATG CAT AAT TGA GGA ACA AGC TGG AAT TAG TGT CAT CGT AAA TGA TCmFlt-HIE50–2ACG TGA TCA TTT ACG ATG ACA CTA ATT CCA CGT TGT TCC TCA ATT ATGmFlt-HIE50–234A1TCG AGA TGG ATG CAT AAT TGA GGA ACA AAA AGG AAT TAG TGT CAT CGT AAA TGA TCmFlt-HIE50–234A2ACG TGA TCA TTT ACG ATG ACA CTA ATT CCT TTT TGT TCC TCA ATT ATG CAT CCA TChEPOHIE-1TCG AGG CCC TAC GTG CTG TCT CAC ACA GCC TGT TCT GAC CTC TCG ACC TAC CGG CChEPOHIE-2TCG AGG CCG GTA GGT CGA GAG GTC AGA ACA GGC TGT GTG AGA CAG CAC GTA GGG CChVEGFHIE-1TCG AGC ACA GTG CAT ACG TGG GCT TCC ACA ChVEGFHIE-2TCG AGT GTG GAA GCC CAC GTA TGC ACT GTG CmFlt-1HIE264.FGAC TAC GCG TCA CGG GTA TCT GGC AGG TTC TAmFlt-1HIE264.RGAC TAC GCG TGA AAC GCT GGA TGG AAA ACA AAmFlt-1HIE100.FGAC TAC GCG TCC GGG ACG ACT TCA GCC TmFlt-1HIE100.RGAC TAC GCG TGG GTG AAA TTA ACT TGA GAC ACT AGA TCHUMKDR 2530.FGGC CAA GAG ATT GAA GCA GAT CHUMKDR 3043.RACT TTC GCG ATG CCA AGA ACT CHUMKDR 2872.FP5′(FAM)-ACT GGT GAT GCT GTC CAA GCG CCG TTT-(TAMRA)p3′HSFLT 2689.RCCC ACT TGC TGG CAT CAT AAG GHSFLT 2228.FCAC CAT ACC TCC TGC GAA ACC THSFLT 2549.FP5′(FAM)-TGG CTG CGA CTC TCT TCT GGC TCC TAT-(TAMRA)p3′ Open table in a new tab Primary cultures of HUVE cells were obtained from Clonetics (San Diego, CA). Cells were maintained in the presence of endothelial growth medium (EGM; Clonetics). EGM consists of endothelial basic medium plus 0.1 ng/ml recombinant human EGF, 10 μg/ml hydrocortisone, 500 μg/ml gentamycin, 500 ng/ml amphotericin B, 12 μg/ml bovine brain extract, and 2% fetal bovine serum. 24 h before exposure to hypoxia, cells were trypsinized and plated on gelatinized 10-cm culture dishes to a density of 8.8 × 103 cells/cm2. Dishes were then floated with preanalyzed gas mixture for 20 min and kept in 0% O2, 5% CO2, and 95% N2 at 37 °C for various durations. Oxygen concentration in the incubators was monitored by a portable oxygen analyzer (Teldyne Brown Engineering, San Leandro, CA). HeLa cells (ATCC number CRL 7923) were maintained in high glucose Dulbecco's modified Eagle's medium (Life Technologies, Inc.) supplemented with 10% fetal bovine serum, 100 units/ml penicillin, and 100 μg/ml streptomycin. Hep3B cells (ATCC number HB-8064) were grown in minimal essential medium with Earle's salts with nonessential amino acids, without glutamine (Life Technologies), complemented with 2 mml-glutamine, 100 units/ml penicillin, and 100 μg/ml streptomycin. For transient expression in Hep3B cells, 0.5 μg of test DNA and 0.5 μg of reference template were added to each well of six-well plates. Cells were at a density of 0.5 × 106/well. Plasmid DNA was prepared by using commercial kits (Qiagen, Santa Clara, CA) and introduced into cells by electroporation with a Gene Pulser (Bio-Rad, Richmond, CA) at 260 V and 960 microfarads following the manufacturer's instructions. Quadruplicate electroporations were pooled and split into six-well plates with 5 ml of medium in each well. Cells were then allowed to recover for 24 h in a 5% CO2, 95% air incubator at 37 °C. Following medium change, triplicate wells were placed in a hypoxic (0% O2, 5% CO2) or in a normoxic (5% CO2, air) incubator for the indicated duration of time. Cell extracts were generated by incubation in 500 μl of 1 × passive lysis buffer solution (dual luciferase assay; Promega) at room temperature for 15 min and frozen at −70 °C. 10 μl of the extracts were analyzed in a luminometer (TD-20e, Turner Designs, Inc., Mountain View, CA) using the reagents provided in the kit. Light production was measured for 15 s, and results were expressed as relative light units (RLU). The mean RLU was corrected with the signals obtained from the reference constructs. The relative luciferase activity (mean ± S.E.) was calculated as luciferase (RLU)/Renilla luciferase (RLU). 1 μg of test DNA and 0.1 μg/well CMV-RLluciferase control vector (dual luciferase assay; Promega) were used for transient transfection of 0.2 × 106 HeLa cells/well in six-well plates. Immediately before the addition of the DNA-liposome complex, cells were washed twice with phosphate-buffered saline. The plasmid DNA mix was then added to a 15-ml Falcon tube containing 0.1 ml of Opti-MEM1 and mixed. 10 μl of Lipofectin (Life Technologies, Inc.) were added to another 15-ml tube containing 0.1 ml of Optim-MEM1 and mixed. Plasmid DNA and Lipofectin were mixed and incubated at room temperature to allow the complex to form. After 15 min, the DNA-liposomal complex was diluted with 0.8 ml of Opti-MEM1 (Life Technologies, Inc.), mixed, and layered gently on top of the cells. Cells were then incubated at 37 °C for 14–16 h. After removal of the DNA-liposome mixture, 3 ml of complemented medium was added 24 h before exposure to hypoxia. For final luciferase assay, cells were lysed in 200-μl passive lysis buffer (Promega). Cell lysates were then transferred into a 1.5-ml microcentrifuge tube and cleared by centrifugation at 10,000 rpm for 1 min at 4 °C. Luciferase andRenilla luciferase activity were measured by mixing 20 μl of extract with 100 μl of luciferase assay buffer and the subsequent addition of 100 μl of Stop and Glow solution, respectively. To generate the luciferase reporter vector containing the mouseFlt-1 promoter, the KpnI-XhoI promoter fragment was ligated into the KpnI and XhoI sites of pGL2-basic plasmid (Promega, Madison, WI). This resulted in the Flt-1-(−2778/+209)-luc construct. A series of 5′-promoter deletions were then generated by double digestion with various restriction enzymes followed by blunt end reaction and religation by standard cloning techniques (25Sambrook J. Fritsch E.F. Maniatis T. Molecular Cloning: A Laboratory Manual.2nd Ed. Cold Spring Harbor Laboratory, Cold Spring Harbor, NY1989Google Scholar). For the Flt-1-(−1464/+209)-luc construct,KpnI and PvuII were used; for Flt-1-(−977/+209)-luc KpnI and NsiI. For Flt-1-(−547/+209)-luc, KpnI and AflII were used, and for the Flt-1-(−202/+209)-luc construct, KpnI andBssHI were used (see Fig. 4 A). The constructs mFlt-HIE50, HIE40, HIE30, HIE50–234A, hEPO HIE, and hVEGF HIE were made by ligation of the corresponding oligonucleotide set into the SalI site 2.8 kb upstream of the SV40 promoter (“enhancer” position) or into the XhoI site (“promoter” position) of the pGL2SV40prom vector (Promega). For oligonucleotide sequences, see Table I. Phosphorylation of 50 pmol of the corresponding oligonucleotide set was carried out in 70 mm Tris-HCl (pH 7.6), 10 mm MgCl2, 5 mm dithiothreitol, 5 mm rATP (Pharmacia Biotech Inc.) and 10 units of T4 polynucleotide kinase (New England Biolabs, Beverly, MA) for 1 h at 37 °C in a total volume of 20 μl. The reaction mixture was heated for 10 min at 65 °C and slowly cooled to room temperature. 0.1 μl of this mix was used in ligation reactions. Constructs mFlt-1HIE264 and mFlt-1HIE100 were generated by PCR (25Sambrook J. Fritsch E.F. Maniatis T. Molecular Cloning: A Laboratory Manual.2nd Ed. Cold Spring Harbor Laboratory, Cold Spring Harbor, NY1989Google Scholar) of pGEM5Flt-1(NcoI) using the oligonucleotides mFlt-1HIE264.F and mFlt-1HIE264.R and oligonucleotides mFlt-1HIE100.F and mFlt-1HIE100.R, respectively. The PCR products were subcloned into pSK to generate pSKmFlt-1HIE264 and pSKmFlt-1HIE100, respectively. The inserts were cut by BamHI and SalI and cloned in “enhancer” position in pGL2prom, as above described. To clone these inserts in “promoter” position in pGL2SV40prom, they were cut withXhoI and SacI and then cloned into theXhoI site of the reporter plasmid. All constructs were analyzed by restriction digestion analysis and partial sequencing. HUVE cells from pooled donors were cultured as described under “Cell Cultures.” Cells were initially expanded for 8–10 days in the presence of EGM. 38 h prior to exposure to hypoxia, cells were split and seeded at a density of 100,000 cells/well in six-well plates in EGM. Immediately prior to hypoxic incubation (0% O2, 5% CO2, 95% N2), cells were washed once, and then 5 ml of assay medium (endothelial basic medium plus 2% fetal bovine serum, 10 μg/ml hydrocortisone, 500 μg/ml gentamycin, 0.5 μg/ml amphotericin-B) was added to each well. After incubations of various duration, cells were harvested by the STAT 60 method (Tel-Test, Inc., Friendswood, TX), and total RNA was prepared according to the manufacturer's recommendations. The RNA was dissolved in 50 μl of H2O, and its concentration was determined by spectrophotometer (A 260). To monitor gene expression, we used real time RT-PCR analysis. This novel approach has been described previously (26Heid C.A. Stevens J. Livak K.J. Williams P.M. Genome Res. 1996; 6: 986-994Crossref PubMed Scopus (5020) Google Scholar, 27Gibson U.E.M. Heid C.A. Williams P.M. Genome Res. 1996; 6: 995-1001Crossref PubMed Scopus (1778) Google Scholar). Briefly, a gene-specific PCR oligonucleotide primer pair defines the “amplicon.” Within the amplicon, an oligonucleotide probe labeled with a reporter fluorescent dye (FAM) at the 5′-end and a quencher fluorescent dye (TAMRA) at the 3′-end are designed. When the probe is intact, the reporter dye emission is quenched. During the extension phase of the PCR cycle, the nucleolytic activity of the DNA polymerase cleaves the hybridization probe and releases the reporter dye from the probe. Fluorescence intensity produced during PCR amplifications is monitored by the sequence detector directly in the reaction tube (“real time”). A computer algorithm compares the amount of reporter dye emission with the quenching dye emission and calculates the threshold cycle number (C T), when signals reach 10 times the standard deviation of the base line. It was demonstrated that the calculated C T values are a quantitative measurement for the mRNA levels of various genes tested (27Gibson U.E.M. Heid C.A. Williams P.M. Genome Res. 1996; 6: 995-1001Crossref PubMed Scopus (1778) Google Scholar). 100 ng of total RNA was added to a 50-μl RT-PCR reaction (PCR-Access, Promega). The reaction master mix was prepared according to the manufacturer's protocol to give final concentrations of 1 × avian myeloblastosis virus/Tfl reaction buffer, 0.2 mm dNTPs, 1.5 mm MgSO4, 0.1 unit/ml avian myeloblastosis virus reverse transcriptase, 0.1 unit/μlTfl DNA polymerase, 250 nm concentration of the primers, and 200 nm concentration of the corresponding probe, as described by Gibson et al. (27Gibson U.E.M. Heid C.A. Williams P.M. Genome Res. 1996; 6: 995-1001Crossref PubMed Scopus (1778) Google Scholar). Primers and probes for real time PCR analysis of Flt-1 andFlk-1/KDR genes were designed by the Oligo version 4.0 program (National Bioscience, Plymouth, MN), according to Heid et al. (26Heid C.A. Stevens J. Livak K.J. Williams P.M. Genome Res. 1996; 6: 986-994Crossref PubMed Scopus (5020) Google Scholar). For sequences of all oligonucleotides used, see TableI. The primers for the humanFlk-1/KDR gene were HUMKDR 2530.F and HUMKDR 3043.R, and the probe was HUMKDR 2872.FP. For Flt-1 analysis, the following primers were used: HSFLT 2689.R and HSFLT 2228.F; the probe was HSFLT 2549.FP. Primers and probes were synthesized at Genentech using conventional nucleic acid synthesis chemistry. The β-actin primer and probe (TaqMan β-actin detection reagents) were purchased from Perkin-Elmer and Applied Biosystems. RT-PCR reactions and the resulting relative increase in reporter fluorescent dye emission were monitored in real time by the 7700 sequence detector (Perkin-Elmer). Signals were analyzed by the sequence detector 1.0 program (Perkin-Elmer). Conditions were as follows: 1 cycle at 48 °C for 45 min, 1 cycle at 94 °C for 2 min, 40 cycles at 94 °C for 30 s, 60 °C for 1 min, 68 °C for 2 min. Data were generated as indicated in the legend to Fig. 1. To determine whether the VEGF receptors are directly regulated by hypoxia in endothelial cells, primary HUVE cells were incubated in hypoxic conditions (0% O2, 5% CO2) and analyzed for VEGF receptor gene expression by real time quantitative RT-PCR analysis (27Gibson U.E.M. Heid C.A. Williams P.M. Genome Res. 1996; 6: 995-1001Crossref PubMed Scopus (1778) Google Scholar). In two independent experiments conducted with different HUVE cell preparations derived from different donors,Flt-1 levels were induced 4.2 ± 0.8-fold after 32 h of growth in hypoxia. Expression levels for Flk-1/KDR were unchanged or weakly down-regulated at the same time point. Time course experiments revealed a 2–3-fold stimulation of Flt-1 after 60 h in hypoxia (Fig.1 A), whereas theFlk-1/KDR levels were moderately down-regulated during the same period (Fig. 1 B). Initial Northern blot experiments yielded essentially similar results (data not shown). To test if the differential regulation of the VEGF receptor genes is due to transcriptional regulatory regions, promoter regions located upstream of both genes were isolated and tested for their ability to respond to hypoxia in fusion constructs with the luciferase gene. To isolate the 5′-flanking region of the murine Flk-1 andFlt-1 genes, a genomic DNA library from 129Sv/ev mice was screened using probes corresponding to the signal peptide sequence of their respective gene products. For each receptor gene, three independent phage clones were isolated and analyzed by restriction mapping. To exclude the possibility of recombination artifacts, genomic DNA from 129/SvJ mice (Stratagene) was analyzed by Southern blot hybridization and showed identical restriction fragments (data not shown). A 3.4-kb KpnI/NcoI Flt-1 and a 2.1-kbPstI Flk-1/KDR were subcloned into pSK vector (Stratagene) and used for further analysis. The nucleotide sequence of the KDR/Flk-1 promoter region (−1829/+148) and the comparison with the human sequence (−780/+148) (28Patterson C. Perrella M.A. Hsieh C.-M. Yoshizumi M. Lee M.-E. Haber E. J. Biol. Chem. 1995; 270: 23111-23118Abstract Full Text Full Text PDF PubMed Scopus (140) Google Scholar) are shown in Fig. 2 A. Mouse Flt-1 promoter sequences (−3181/+276) and the comparison with the available human gene sequence (−1195/+276) (29Morishita K. Johnson D.E. Williams L.T. J. Biol. Chem. 1995; 270: 27948-27953Abstract Full Text Full Text PDF PubMed Scopus (106) Google Scholar,30Ikeda T. Wakiya K. Shibuya M. Growth Factors. 1996; 13: 1151-1162Crossref Scopus (22) Google Scholar), are shown in Fig. 2 B. The sequence comparison between the human and mouse Flt-1 genes revealed 78% similarity in a 1.5-kb promoter region (Fig. 2 A). Flk-1/KDRexhibited a 60% similarity in a 0.9-kb promoter region. Interestingly, comparison of the coding regions between the human and mouse genes reveals a 75% similarity for Flt-1 (31Shibuya M. Yamaguchi S. Yamane A. Ikeda T. Tojo A. Matsushime H. Sato M. Oncogene. 1990; 8: 519-527Google Scholar, 32Finnerty H. Kelleher K. Morris G.E. Bean K. Merberg D.M. Kriz R. Morris J.C. Sookdeo H. Turner K.J. Wood C.R. Oncogene. 1993; 8: 2293-2298PubMed Google Scholar) and 85% for Flk-1/KDR (33Matthews W. Jordan C.T. Gavin M. Jenkins N.A. Copeland N.G. Lemischka I.R. Proc. Natl. Acad. Sci. U. S. A. 1991; 88: 9026-9030Crossref PubMed Scopus (450) Google Scholar, 34Terman B.I. Carrion M.E. Kovacs E. Rasmussen B.A. Eddy R.L. Shows T.B. Oncogene. 1991; 6: 1677-1683PubMed Google Scholar). T