Editorial Open Access
Copyright ©2006 Baishideng Publishing Group Co., Limited. All rights reserved.
World J Gastroenterol. Aug 21, 2006; 12(31): 4937-4942
Published online Aug 21, 2006. doi: 10.3748/wjg.v12.i31.4937
Post-transcriptional regulation of vascular endothelial growth factor: Implications for tumor angiogenesis
Peter S Yoo, Abby L Mulkeen, Charles H Cha, Gastrointestinal Surgery and Surgical Oncology, Yale School of Medicine, New Haven, Connecticut, United States
Author contributions: All authors contributed equally to the work.
Correspondence to: Charles H Cha, MD, Department of Surgery, Yale School of Medicine, 330 Cedar Street, LH 118, New Haven, CT 06520-8062, United States. charles.cha@yale.edu
Telephone: +1-203-7852380 Fax: +1-203-9373845
Received: December 1, 2005
Revised: January 5, 2006
Accepted: January 14, 2006
Published online: August 21, 2006

Abstract

Vascular endothelial growth factor (VEGF) is a potent secreted mitogen critical for physiologic and tumor angiogenesis. Regulation of VEGF occurs at several levels, including transcription, mRNA stabilization, translation, and differential cellular localization of various isoforms. Recent advances in our understanding of post-transcriptional regulation of VEGF include identification of the stabilizing mRNA binding protein, HuR, and the discovery of internal ribosomal entry sites in the 5'UTR of the VEGF mRNA. Monoclonal anti-VEGF antibody was recently approved for use in humans, but suffers from the need for high systemic doses. RNA interference (RNAi) technology is being used in vitro and in animal models with promising results. Here, we review the literature on post-transcriptional regulation of VEGF and describe recent progress in targeting these mechanisms for therapeutic benefit.

Key Words: Vascular endothelial growth factor, Vascular endothelial growth hormone, Post-transcriptional regulation, mRNA stability, HuR, ELAV1, Internal ribosomal entry, IRES, siRNA, RNAi, Bevacizumab



INTRODUCTION

The growth of any tissue in vivo relies on the adequacy of its vascular supply. Moreover, the notion of tumor angiogenesis, first advanced by Folkman in the early 1970s, posits that the growth of any tumor beyond 1-2 cubic centimeters depends on the growth of new blood supply[1]. In cases of insufficient vasculature, tissues become depleted of oxygen and nutrients, leading to the secretion of angiogenic factors[2]. These factors, which include hormones, interleukins, insulin, and growth factors, spur the ingrowth of new blood vessels. Vascular endothelial growth factor (VEGF) is one such key regulator of angiogenesis. It is a secreted mitogen and the most potent angiogenic factor, specific to endothelial cells and highly expressed in areas of active angiogenesis such as solid tumors.

The upregulation of VEGF expression in response to hypoxia plays a crucial role in tumor angiogenesis. Though it is not itself an oncogene, VEGF is upregulated in tumorigenesis and is important in blood vessel formation in solid tumors[3]. VEGF levels correlate with tumor progression and invasion, and a high VEGF level in colorectal carcinoma has been found to be an independent prognostic factor for long-term survival[4,5].

Hypoxic induction of VEGF appears to occur both by transcriptional activation and through stabilization of VEGF mRNA, which is otherwise labile in normoxic conditions. Transcriptional activation of VEGF relies largely on binding to the hypoxia inducible factor-1 (HIF-1), a heterodimeric basic helix-loop-helix protein that activates transcription of the human erythropoietin gene in hypoxic cells[6]. HIF-1 binds to a sequence in the 5'-flanking region of the VEGF gene called the hypoxic response element (HRE)[7-9]. Several other molecules have been implicated in the transcriptional control of VEGF. Sp1 stimulates transcription by binding to G/C-rich boxes present on the VEGF promoter[10]. AP-1 is a dimeric transcription factor of the leucine zipper family that is composed of jun/jun or jun/fos subunits. Hypoxia, oxidative stress, and cytokines may increase VEGF expression through the synthesis of jun and fos proteins and increased AP-1 binding activity. A number of other transcription factors also contribute to VEGF induction and regulation, including Stat-3[11].

Alternative splicing

In humans there are five alternatively spliced isoforms of VEGF, each is named for the number of amino acids along its length (VEGF 206, -189, -165, -145, 121). While varying amounts of each VEGF isoform mRNA can be generated to produce certain or all isoforms of VEGF[12], each is predicted to have a characteristic extracellular localization based on biochemical differences. The larger isoforms bind neuropilin, matrix, and cell surface heparin proteoglycans, and are thought to act locally. The smaller isoforms do not display the heparin proteoglycan binding region and may diffuse to sites distant from the site of synthesis[13,14].

Post-transcriptional Regulation

A body of evidence is growing rapidly to demonstrate that post-transcriptional regulation of VEGF is critically important in the fine-tuned response required for both physiologic and malignant expression. The importance of post-transcriptional regulation of VEGF is based upon several key observations about the intrinsic nature of the VEGF mRNA. First, the mRNA is labile under conditions of normal oxygen tension[15]. Second, the 5' UTR of the VEGF mRNA possesses features that make efficient ribosome scanning and initiation of translation unwieldy.

Lability

Whereas the average half-life of eukaryotic mRNAs is 10-12 h, the half-life of VEGF mRNA is less than one hour[16]. In 1995, Ikeda and colleagues first reported that although VEGF levels were elevated in cells cultured in hypoxic conditions, transcriptional activation alone could not account for the increase in VEGF mRNA levels. Nuclear run-on assays demonstrated upregulation of VEGF transcription under hypoxia that was apparent after three hours, and persisted after fifteen hours of incubation. VEGF mRNA levels were 8-10 times higher than baseline, and this persistent elevation of VEGF could only be explained by increased stability of the mRNA. They concluded that hypoxia could lengthen the half-life of VEGF mRNA by 2-3 fold[7].

The effect of stabilization is mediated by the RNA-binding protein HuR. Steitz and colleagues first established the association between mRNA degradation and HuR, a member of the Elav family of proteins found in Drosophila[17]. Beginning with the observation that some mRNAs are targeted for rapid degradation by the presence of AU-rich elements (AREs) in the 3' UTRs, it was determined that HuR exhibits affinity for AREs, and levels of HuR correlate with in vitro mRNA degradation[18].

In VEGF, it was observed that the increase in mRNA stability coincided with the binding of a protein to an ARE, forming an RNA-protein complex in a hypoxia-inducible fashion[16]. Using a tumor cell line lacking the wild type von Hippel-Lindau tumor suppressor gene and in which VEGF mRNA is constitutively stabilized, this RNA-protein complex was found to be constitutively elevated[19]. The protein was later identified as the HuR protein[15]. Inhibition of HuR expression by antisense sequences was found to inhibit the hypoxic stabilization of VEGF mRNA, demonstrating its critical role in post-transcriptional stabilization of VEGF expression. However, total cellular steady-state HuR was not altered by hypoxia, raising the possibility that HuR is a component of a hypoxia-inducible complex whose other components are regulated by oxygen tension[15].

HuR probably has a function as a nuclear-cytoplasmic shuttle as well. It is predominantly localized to the nucleus, where it likely binds newly-transcribed VEGF mRNA and transports it to the cytoplasm and protects it from the ARE-mediated degradation pathways[20].

Recent studies have suggested HuR and VEGF mRNA localize to the cytoplasm in response to cellular stress[21]. Upon hypoxic stress, HuR is found to be localized to the cytosol and bound to target mRNAs, preventing their decay. New experiments using double labeling immunofluorescence show that VEGF and HuR colocalize to the nucleus during hypoxia[22,23]. The nuclear role of VEGF mRNA, if any is unclear. Further studies to more clearly delineate the intracellular localization of VEGF mRNA and HuR are warranted.

The mechanism by which HuR protects VEGF from degradation has been further investigated. One hypothesis is that RNA binding proteins confer stability by binding directly to destabilizing sequences and making them unavailable to endonucleases. Goldberg and colleagues suggest that RNA-stabilizing factors may localize to a distinct binding site and thereby change the secondary or tertiary structure of the RNA, making a specific site unavailable to endonucleolysis. In support of this hypothesis, a 40 bp region adjacent to the HuR-binding site in the VEGF stability region has been identified that is susceptible to ribonucleases in the absence of HuR[24].

HuR's importance in carcinogenesis is underlined by a report Lopez de Silanes and colleagues[25]. Using a nude mouse model and the RKO colorectal cancer cell line, they reported that tumors modified to overexpress HuR grew significantly larger than controls, and conversely, tumors modified to repress HuR expression grew smaller. Whether this effect was brought about through stabilization of VEGF mRNA is yet unknown.

IRES

Internal ribosomal entry sites (IRESs) were first discovered in picornaviruses in which they initiate translation of naturally uncapped viral mRNA. Eukaryotic mRNA, however, possesses a 7-methylguanosine cap in the 5'UTR that is critical in the canonical model of translation. In this model, the eukaryotic initiation factors (eIFs) recognize and bind the 5' cap, and unwind the secondary structure of the 5'UTR, thereby making it sterically feasible for ribosomes to scan through the 5' UTR and translate the mRNA beginning at the AUG start codon[26].

However, perhaps 3%-5% of all cellular mRNAs are translated by a mechanism independent of the 5' cap structure[27]. Most of these mRNAs are likely to contain an internal ribosome-entry site (IRES) in the 5' untranslated region (UTR), as the ability to maintain efficient translation without utilizing the 5' cap-dependent mechanism relies on the presence of one or several IRESs. IRESs are typically found in mRNAs that possess unusually long 5'UTRs, greater than -300 nt, which consequently have significant secondary structure[28]. In eukaryotes, IRES-mediated translation bypasses the canonical initiation step by directly recruiting ribosomes to the start codon without the need for the 5' mRNA cap structure or the eukaryotic initiation factor (eIF) complex[28,29].

The eIF-4F complex is one of the key mediators of ribosome recruitment in cap-dependent translation and the eIF-4E subunit of this complex can be rate-limiting[2]. Under normal cellular conditions, eIF-4E is bound to its binding protein 4E-BP1 and translation is limited by the availability of the unbound fraction of eIF-4E[30]. When 4E-BP1 is phosphorylated, eIF-4E is released and becomes available to bind mRNA and participate in the eIF-4F complex. This phosphorylation is instigated by the downstream effects of various mitogenic pathways. Hypoxia, however, has the opposite effect, increasing the binding of eIF-4E with 4E-BP1, which may mediate the global decrease in protein synthesis seen in stressed cells[31].

Several observations about the 5' UTR of VEGF mRNA led to the hypothesis that it may contain IRES elements. First, cap-dependent translation of VEGF is cumbersome because the 5' UTR (1038 bp in humans) is much longer than typical eukaryotic 5' UTRs (-300 nt) and does not allow efficient ribosomal scanning. Second, it has a high G+C content, predisposing it to form stable secondary structures. Third, the 5'UTR contains a short open reading frame with in-frame initiation and termination codons[32]. Furthermore, in order to maximize its effect, translation of VEGF must be upregulated during periods of stress such as cellular hypoxia, when cap-dependent protein synthesis is globally inhibited.

In 1998, VEGF mRNA was shown to possess two IRES sites (IRES A and B) in the 5' UTR[32-34]. IRES A is contained in a 293nt segment just upstream from the AUG codon and is believed to control translation initiated at the AUG codon[35]. IRES B is contained in the early portion of the 5' UTR[33]. These two sites appear to bind different proteins, and the cellular milieus in which each functions optimally are not well understood.

IRES-mediated translation in vivo and in vitro yields a distinct isoform of VEGF, called “Large VEGF” or “L-VEGF.” Whereas initiation mediated by the 5' cap and IRES A is at the AUG start codon it shares with cap-dependent translation, the initiation of the L-VEGF isoform occurs at an upstream CUG start codon (nt 499) within the 5'UTR. The L-VEGF isoform contains an N-terminal extension of 180 a.a., which is cleaved to yield a C-terminal peptide that is nearly identical to VEGF 189. This N-terminal extension is highly conserved through human, bovine, and murine cells, but whether it has any function is unknown[36].

Inquiry into the control of IRES A has uncovered a complex model of translational regulation. The five classic isoforms (excluding L-VEGF) are the translation products of five distinct mature mRNAs; each of these mature mRNAs is the product of alternative splicing of the transcription product or pre-mRNA. Bornes and colleagues have reported that differences in exon content of the differentially spliced mRNA can lead to preferential activity of one IRES or the other[35].

Though it is becoming clear that the IRES mechanism in VEGF translation is a plausible pathway in normal physiology, it remains debatable whether this pathway plays a significant role in tumor angiogenesis. The controversy relies on the observation that there is a correlation between eIF-4E and protein synthesis in malignancy. This correlation has been demonstrated in breast carcinomas, head and neck squamous cell carcinomas, soft tissue sarcomas, and colon carcinoma[37-40], and is bolstered by the observation that cells that over-express eIF-4E increase secretion of VEGF by greater than 100-fold[41]. Therefore, it is conceivable that VEGF production in transformed cells relies on the cap-dependent pathway as well, though this has not been definitively demonstrated.

Whether hypoxia is a sufficient stimulus to induce IRES activity, or if other mitogenic stimuli can trigger internal ribosomal entry remains unknown. In a study comparing the IRES elements of VEGF, hypoglycemia was seen to significantly activate IRES activity[42]. In both transformed and benign cells, the balance between IRES and cap-dependent translation in VEGF requires further investigation.

Deranged expression of the c-myc oncogene is associated with many human malignancies, as its over-expression promotes cell growth and angiogenesis[43]. This effect is in part due to the fact that myc over-expression increases secretion of VEGF[44]. The mechanism behind this increase was recently described by Mezquita and colleagues[45]. They demonstrated that myc interacts with VEGF mRNA to upregulate initiation of translation capable of bringing about a 10-fold increase in VEGF production. They observed no increase in eIF activity and attribute the translational upregulation to VEGF-specific increase of translation initiation, either through c-myc or VEGF directly.

Implications for Tumor Angiogenesis

In 1993, Kim and colleagues reported that anti-VEGF antibodies had a strong inhibitory effect on the growth of several tumor types in nude mice. Since then, anti-VEGF therapies to target angiogenesis have gained further attention. Investigators have reported inhibited tumor growth with other anti-VEGF applications including anti-sense nucleotides, anti-receptor antibodies, soluble VEGF receptors, and a retrovirus-delivered Flk-1 mutant.[46-49] A recent phase three FDA trial demonstrated improved survival when patients with metastatic colorectal cancer were treated with a humanized monoclonal anti-VEGF antibody (bevacizumab) in combination with standard chemotherapy (irinotecan, 5-FU, and leucovorin), resulting in FDA approval. The FDA has also approved its use in non-small cell lung cancer. Bevacizumab is being tested in several other tumor models with and without adjunctive traditional chemotherapy agents.

Anti-angiogenic therapies have targeted the normal endothelial cell's ability to respond to angiogenic factors, either by primarily inhibiting endothelial cell proliferation or by using antibodies to prevent angiogenic factors from activating endothelial cells[50]. However, some investigators have observed resistance to this therapy due to the ability of tumor cells to upregulate the expression of VEGF and thereby negate the anti-angiogenic effect[51]. There is further evidence suggesting that high tumor levels of VEGF may promote tumor survival[52].

One potential anti-angiogenic strategy is to silence gene expression of angiogenic factors at the mRNA level.While antisense oligonucleotides have been used for this strategy, the high concentration required and non-specific effects have limited this approach. First described in C. elegans, RNA interference (RNAi) is a process in which double stranded RNA (dsRNA) is processed by the enzyme Dicer, resulting in short interfering RNAs (siRNA) 21-25 nucleotides in length[53,54]. These siRNA molecules are incorporated into an RNA-induced silencing complex (RISC) that binds complementary target mRNAs, leading to their degradation[55,56].

There is great potential for using RNAi technology in therapeutic applications that target critical signaling pathways involved in cancer, due to its high specificity for targeted genes and its potency at low concentrations. However, relatively few studies have examined the use of RNAi technology to suppress VEGF production.

In 2003, three reports proved the concept. Filleur and colleagues used a rat fibrosarcoma model and systemically delivered siRNA to achieve a 66% decrease in tumor size when compared to controls[57]. Reich and colleagues used direct injection of anti-VEGF siRNA to decrease choroidal neovascularization in laser-induced retinal injury in the rat[58]. A plasmid-based siRNA delivery system driven by a Pol III promoter was used by Zhang and colleagues to show isoform-specific knockdown of VEGF[59].

These three experiments provided the conceptual basis for two further experiments to demonstrate the efficacy of RNAi inhibition of VEGF. In 2005, Yoon and colleagues injected HT1080 fibrosarcoma cells stably transfected with anti-vegf siRNA into the subcutaneous tissue of rats. They reported substantial decrease in tumor volume and vessel density in resected specimens[60]. In the same year, Kwon and colleagues described a bi-level approach to silencing VEGF. In their experiment, they combined zinc finger-mediated repression of transcription with RNAi technology to achieve knockdown of VEGF of over 90%[61].

The critical problem with RNAi-based therapy is one of delivery. Whereas in vitro transfection of cells is easily accomplished using lipid-based solution, in vivo targeting of the siRNA molecules to cancer cells is far more challenging. Systemic delivery of naked siRNA is problematic as the drugs localize poorly and the degree of dilution is often unacceptable[62]. Though synthetic vectors offer several advantages such as their relative safety and the ease of incorporating target-specific ligands, there are significant barriers to their effective use that have not yet been overcome[63]. Viral vectors may offer specific advantages as well[64].

CONCLUSION

Although much has been learned about VEGF since its discovery in 1989, much more remains to be understood. First, the relevance of hypoxic stabilization of VEGF mRNA must be established in the settings of both neoplasia and normal physiologic conditions. Does our current understanding of hypoxic stabilization provide a sufficient explanation with regard to tumor angiogenesis Are there other mitogenic factors that can overwhelm the cell's innate response to hypoxia Second, the balance between cap-dependent and IRES-mediated translation must be elucidated. What are the factors that mitigate the switch from cap-dependent to IRES-mediated translation Finally, it is clear that post-transcriptional regulation of VEGF has profound clinical implications with regard to targeting tumor angiogenesis. Will RNAi technology prove to be an efficacious therapeutic modality If so, how will we deliver the drug to achieve a safe, durable, and clinically significant outcome It will be a challenge for coming years to determine how best to exploit these molecular mechanisms for our patients' benefit.

Footnotes

S- Editor Wang J L- Editor Worthley DL E- Editor Bai SH

References
1.  Folkman J. Tumor angiogenesis: therapeutic implications. N Engl J Med. 1971;285:1182-1186.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 5115]  [Cited by in F6Publishing: 5647]  [Article Influence: 106.5]  [Reference Citation Analysis (0)]
2.  Hiremath LS, Webb NR, Rhoads RE. Immunological detection of the messenger RNA cap-binding protein. J Biol Chem. 1985;260:7843-7849.  [PubMed]  [DOI]  [Cited in This Article: ]
3.  Plate KH, Breier G, Weich HA, Risau W. Vascular endothelial growth factor is a potential tumour angiogenesis factor in human gliomas in vivo. Nature. 1992;359:845-848.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1572]  [Cited by in F6Publishing: 1568]  [Article Influence: 49.0]  [Reference Citation Analysis (0)]
4.  Cascinu S, Staccioli MP, Gasparini G, Giordani P, Catalano V, Ghiselli R, Rossi C, Baldelli AM, Graziano F, Saba V. Expression of vascular endothelial growth factor can predict event-free survival in stage II colon cancer. Clin Cancer Res. 2000;6:2803-2807.  [PubMed]  [DOI]  [Cited in This Article: ]
5.  Poon RT, Ng IO, Lau C, Zhu LX, Yu WC, Lo CM, Fan ST, Wong J. Serum vascular endothelial growth factor predicts venous invasion in hepatocellular carcinoma: a prospective study. Ann Surg. 2001;233:227-235.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 192]  [Cited by in F6Publishing: 197]  [Article Influence: 8.6]  [Reference Citation Analysis (0)]
6.  Forsythe JA, Jiang BH, Iyer NV, Agani F, Leung SW, Koos RD, Semenza GL. Activation of vascular endothelial growth factor gene transcription by hypoxia-inducible factor 1. Mol Cell Biol. 1996;16:4604-4613.  [PubMed]  [DOI]  [Cited in This Article: ]
7.  Ikeda E, Achen MG, Breier G, Risau W. Hypoxia-induced transcriptional activation and increased mRNA stability of vascular endothelial growth factor in C6 glioma cells. J Biol Chem. 1995;270:19761-19766.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 382]  [Cited by in F6Publishing: 408]  [Article Influence: 14.1]  [Reference Citation Analysis (0)]
8.  Shima DT, Deutsch U, D'Amore PA. Hypoxic induction of vascular endothelial growth factor (VEGF) in human epithelial cells is mediated by increases in mRNA stability. FEBS Lett. 1995;370:203-208.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 206]  [Cited by in F6Publishing: 223]  [Article Influence: 7.7]  [Reference Citation Analysis (0)]
9.  Laughner E, Taghavi P, Chiles K, Mahon PC, Semenza GL. HER2 (neu) signaling increases the rate of hypoxia-inducible factor 1alpha (HIF-1alpha) synthesis: novel mechanism for HIF-1-mediated vascular endothelial growth factor expression. Mol Cell Biol. 2001;21:3995-4004.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 945]  [Cited by in F6Publishing: 938]  [Article Influence: 40.8]  [Reference Citation Analysis (0)]
10.  Shi Q, Le X, Abbruzzese JL, Peng Z, Qian CN, Tang H, Xiong Q, Wang B, Li XC, Xie K. Constitutive Sp1 activity is essential for differential constitutive expression of vascular endothelial growth factor in human pancreatic adenocarcinoma. Cancer Res. 2001;61:4143-4154.  [PubMed]  [DOI]  [Cited in This Article: ]
11.  Niu G, Wright KL, Huang M, Song L, Haura E, Turkson J, Zhang S, Wang T, Sinibaldi D, Coppola D. Constitutive Stat3 activity up-regulates VEGF expression and tumor angiogenesis. Oncogene. 2002;21:2000-2008.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 877]  [Cited by in F6Publishing: 912]  [Article Influence: 41.5]  [Reference Citation Analysis (0)]
12.  Ng YS, Rohan R, Sunday ME, Demello DE, D'Amore PA. Differential expression of VEGF isoforms in mouse during development and in the adult. Dev Dyn. 2001;220:112-121.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in F6Publishing: 4]  [Reference Citation Analysis (0)]
13.  Park JE, Keller GA, Ferrara N. The vascular endothelial growth factor (VEGF) isoforms: differential deposition into the subepithelial extracellular matrix and bioactivity of extracellular matrix-bound VEGF. Mol Biol Cell. 1993;4:1317-1326.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 816]  [Cited by in F6Publishing: 789]  [Article Influence: 25.5]  [Reference Citation Analysis (0)]
14.  Neufeld G, Cohen T, Gengrinovitch S, Poltorak Z. Vascular endothelial growth factor (VEGF) and its receptors. FASEB J. 1999;13:9-22.  [PubMed]  [DOI]  [Cited in This Article: ]
15.  Levy NS, Chung S, Furneaux H, Levy AP. Hypoxic stabilization of vascular endothelial growth factor mRNA by the RNA-binding protein HuR. J Biol Chem. 1998;273:6417-6423.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 489]  [Cited by in F6Publishing: 515]  [Article Influence: 19.8]  [Reference Citation Analysis (0)]
16.  Levy AP, Levy NS, Goldberg MA. Post-transcriptional regulation of vascular endothelial growth factor by hypoxia. J Biol Chem. 1996;271:2746-2753.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 485]  [Cited by in F6Publishing: 497]  [Article Influence: 17.8]  [Reference Citation Analysis (0)]
17.  Robinow S, Campos AR, Yao KM, White K. The elav gene product of Drosophila, required in neurons, has three RNP consensus motifs. Science. 1988;242:1570-1572.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 223]  [Cited by in F6Publishing: 253]  [Article Influence: 7.0]  [Reference Citation Analysis (0)]
18.  Fan XC, Steitz JA. Overexpression of HuR, a nuclear-cytoplasmic shuttling protein, increases the in vivo stability of ARE-containing mRNAs. EMBO J. 1998;17:3448-3460.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 645]  [Cited by in F6Publishing: 701]  [Article Influence: 27.0]  [Reference Citation Analysis (0)]
19.  Levy AP, Levy NS, Goldberg MA. Hypoxia-inducible protein binding to vascular endothelial growth factor mRNA and its modulation by the von Hippel-Lindau protein. J Biol Chem. 1996;271:25492-25497.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 157]  [Cited by in F6Publishing: 146]  [Article Influence: 5.2]  [Reference Citation Analysis (0)]
20.  Fan XC, Steitz JA. HNS, a nuclear-cytoplasmic shuttling sequence in HuR. Proc Natl Acad Sci USA. 1998;95:15293-15298.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 353]  [Cited by in F6Publishing: 386]  [Article Influence: 14.8]  [Reference Citation Analysis (0)]
21.  Brennan CM, Steitz JA. HuR and mRNA stability. Cell Mol Life Sci. 2001;58:266-277.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 805]  [Cited by in F6Publishing: 837]  [Article Influence: 36.4]  [Reference Citation Analysis (0)]
22.  Leung DW, Cachianes G, Kuang WJ, Goeddel DV, Ferrara N. Vascular endothelial growth factor is a secreted angiogenic mitogen. Science. 1989;246:1306-1309.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3536]  [Cited by in F6Publishing: 3491]  [Article Influence: 99.7]  [Reference Citation Analysis (0)]
23.  Lejbkowicz F, Goldberg-Cohen I, Levy AP. New horizons for VEGF. Is there a role for nuclear localization. Acta Histochem. 2005;106:405-411.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 26]  [Cited by in F6Publishing: 26]  [Article Influence: 1.4]  [Reference Citation Analysis (0)]
24.  Goldberg-Cohen I, Furneauxb H, Levy AP. A 40-bp RNA element that mediates stabilization of vascular endothelial growth factor mRNA by HuR. J Biol Chem. 2002;277:13635-13640.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 98]  [Cited by in F6Publishing: 101]  [Article Influence: 4.6]  [Reference Citation Analysis (0)]
25.  López de Silanes I, Fan J, Yang X, Zonderman AB, Potapova O, Pizer ES, Gorospe M. Role of the RNA-binding protein HuR in colon carcinogenesis. Oncogene. 2003;22:7146-7154.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 208]  [Cited by in F6Publishing: 232]  [Article Influence: 11.0]  [Reference Citation Analysis (0)]
26.  Pain VM. Initiation of protein synthesis in eukaryotic cells. Eur J Biochem. 1996;236:747-771.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 535]  [Cited by in F6Publishing: 574]  [Article Influence: 20.5]  [Reference Citation Analysis (0)]
27.  Johannes G, Carter MS, Eisen MB, Brown PO, Sarnow P. Identification of eukaryotic mRNAs that are translated at reduced cap binding complex eIF4F concentrations using a cDNA microarray. Proc Natl Acad Sci USA. 1999;96:13118-13123.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 284]  [Cited by in F6Publishing: 309]  [Article Influence: 12.4]  [Reference Citation Analysis (0)]
28.  Hellen CU, Sarnow P. Internal ribosome entry sites in eukaryotic mRNA molecules. Genes Dev. 2001;15:1593-1612.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 748]  [Cited by in F6Publishing: 736]  [Article Influence: 32.0]  [Reference Citation Analysis (0)]
29.  Johannes G, Sarnow P. Cap-independent polysomal association of natural mRNAs encoding c-myc, BiP, and eIF4G conferred by internal ribosome entry sites. RNA. 1998;4:1500-1513.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 197]  [Cited by in F6Publishing: 207]  [Article Influence: 8.0]  [Reference Citation Analysis (0)]
30.  Sonenberg N, Gingras AC. The mRNA 5' cap-binding protein eIF4E and control of cell growth. Curr Opin Cell Biol. 1998;10:268-275.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 458]  [Cited by in F6Publishing: 460]  [Article Influence: 17.7]  [Reference Citation Analysis (0)]
31.  Tinton SA, Buc-Calderon PM. Hypoxia increases the association of 4E-binding protein 1 with the initiation factor 4E in isolated rat hepatocytes. FEBS Lett. 1999;446:55-59.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 47]  [Cited by in F6Publishing: 49]  [Article Influence: 2.0]  [Reference Citation Analysis (0)]
32.  Stein I, Itin A, Einat P, Skaliter R, Grossman Z, Keshet E. Translation of vascular endothelial growth factor mRNA by internal ribosome entry: implications for translation under hypoxia. Mol Cell Biol. 1998;18:3112-3119.  [PubMed]  [DOI]  [Cited in This Article: ]
33.  Huez I, Créancier L, Audigier S, Gensac MC, Prats AC, Prats H. Two independent internal ribosome entry sites are involved in translation initiation of vascular endothelial growth factor mRNA. Mol Cell Biol. 1998;18:6178-6190.  [PubMed]  [DOI]  [Cited in This Article: ]
34.  Miller DL, Dibbens JA, Damert A, Risau W, Vadas MA, Goodall GJ. The vascular endothelial growth factor mRNA contains an internal ribosome entry site. FEBS Lett. 1998;434:417-420.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 62]  [Cited by in F6Publishing: 71]  [Article Influence: 2.7]  [Reference Citation Analysis (0)]
35.  Bornes S, Boulard M, Hieblot C, Zanibellato C, Iacovoni JS, Prats H, Touriol C. Control of the vascular endothelial growth factor internal ribosome entry site (IRES) activity and translation initiation by alternatively spliced coding sequences. J Biol Chem. 2004;279:18717-18726.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 48]  [Cited by in F6Publishing: 51]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
36.  Huez I, Bornes S, Bresson D, Créancier L, Prats H. New vascular endothelial growth factor isoform generated by internal ribosome entry site-driven CUG translation initiation. Mol Endocrinol. 2001;15:2197-2210.  [PubMed]  [DOI]  [Cited in This Article: ]
37.  Kerekatte V, Smiley K, Hu B, Smith A, Gelder F, De Benedetti A. The proto-oncogene/translation factor eIF4E: a survey of its expression in breast carcinomas. Int J Cancer. 1995;64:27-31.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 140]  [Cited by in F6Publishing: 148]  [Article Influence: 5.1]  [Reference Citation Analysis (0)]
38.  Nathan CO, Liu L, Li BD, Abreo FW, Nandy I, De Benedetti A. Detection of the proto-oncogene eIF4E in surgical margins may predict recurrence in head and neck cancer. Oncogene. 1997;15:579-584.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 139]  [Cited by in F6Publishing: 148]  [Article Influence: 5.5]  [Reference Citation Analysis (0)]
39.  McClusky DR, Chu Q, Yu H, Debenedetti A, Johnson LW, Meschonat C, Turnage R, McDonald JC, Abreo F, Li BD. A prospective trial on initiation factor 4E (eIF4E) overexpression and cancer recurrence in node-positive breast cancer. Ann Surg. 2005;242:584-590; discussion 590-592.  [PubMed]  [DOI]  [Cited in This Article: ]
40.  Vazquez SH, Byrnes KW, Chu Q, Cole P, Dunn G, Werner A, Grimes F, Stratton M, Sittig K, Li B. Eukaryotic initiation factor 4E (eIF4E) expression in malignant versus inflammatory colon tissue. Annals Of Surgical Oncology. 2005;12:S90-S91.  [PubMed]  [DOI]  [Cited in This Article: ]
41.  Kevil CG, De Benedetti A, Payne DK, Coe LL, Laroux FS, Alexander JS. Translational regulation of vascular permeability factor by eukaryotic initiation factor 4E: implications for tumor angiogenesis. Int J Cancer. 1996;65:785-790.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in F6Publishing: 1]  [Reference Citation Analysis (0)]
42.  Wong ET, Ngoi SM, Lee CG. Improved co-expression of multiple genes in vectors containing internal ribosome entry sites (IRESes) from human genes. Gene Ther. 2002;9:337-344.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 48]  [Cited by in F6Publishing: 48]  [Article Influence: 2.2]  [Reference Citation Analysis (0)]
43.  Nesbit CE, Tersak JM, Prochownik EV. MYC oncogenes and human neoplastic disease. Oncogene. 1999;18:3004-3016.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 855]  [Cited by in F6Publishing: 870]  [Article Influence: 34.8]  [Reference Citation Analysis (0)]
44.  Pelengaris S, Littlewood T, Khan M, Elia G, Evan G. Reversible activation of c-Myc in skin: induction of a complex neoplastic phenotype by a single oncogenic lesion. Mol Cell. 1999;3:565-577.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 372]  [Cited by in F6Publishing: 392]  [Article Influence: 15.7]  [Reference Citation Analysis (0)]
45.  Mezquita P, Parghi SS, Brandvold KA, Ruddell A. Myc regulates VEGF production in B cells by stimulating initiation of VEGF mRNA translation. Oncogene. 2005;24:889-901.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 52]  [Cited by in F6Publishing: 55]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
46.  Gerber HP, Kowalski J, Sherman D, Eberhard DA, Ferrara N. Complete inhibition of rhabdomyosarcoma xenograft growth and neovascularization requires blockade of both tumor and host vascular endothelial growth factor. Cancer Res. 2000;60:6253-6258.  [PubMed]  [DOI]  [Cited in This Article: ]
47.  Goldman CK, Kendall RL, Cabrera G, Soroceanu L, Heike Y, Gillespie GY, Siegal GP, Mao X, Bett AJ, Huckle WR. Paracrine expression of a native soluble vascular endothelial growth factor receptor inhibits tumor growth, metastasis, and mortality rate. Proc Natl Acad Sci USA. 1998;95:8795-8800.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 309]  [Cited by in F6Publishing: 323]  [Article Influence: 12.4]  [Reference Citation Analysis (0)]
48.  Millauer B, Shawver LK, Plate KH, Risau W, Ullrich A. Glioblastoma growth inhibited in vivo by a dominant-negative Flk-1 mutant. Nature. 1994;367:576-579.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 857]  [Cited by in F6Publishing: 926]  [Article Influence: 30.9]  [Reference Citation Analysis (0)]
49.  Prewett M, Huber J, Li Y, Santiago A, O'Connor W, King K, Overholser J, Hooper A, Pytowski B, Witte L. Antivascular endothelial growth factor receptor (fetal liver kinase 1) monoclonal antibody inhibits tumor angiogenesis and growth of several mouse and human tumors. Cancer Res. 1999;59:5209-5218.  [PubMed]  [DOI]  [Cited in This Article: ]
50.  Liekens S, De Clercq E, Neyts J. Angiogenesis: regulators and clinical applications. Biochem Pharmacol. 2001;61:253-270.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 499]  [Cited by in F6Publishing: 518]  [Article Influence: 22.5]  [Reference Citation Analysis (0)]
51.  Viloria-Petit A, Crombet T, Jothy S, Hicklin D, Bohlen P, Schlaeppi JM, Rak J, Kerbel RS. Acquired resistance to the antitumor effect of epidermal growth factor receptor-blocking antibodies in vivo: a role for altered tumor angiogenesis. Cancer Res. 2001;61:5090-5101.  [PubMed]  [DOI]  [Cited in This Article: ]
52.  Wong AK, Alfert M, Castrillon DH, Shen Q, Holash J, Yancopoulos GD, Chin L. Excessive tumor-elaborated VEGF and its neutralization define a lethal paraneoplastic syndrome. Proc Natl Acad Sci USA. 2001;98:7481-7486.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 53]  [Cited by in F6Publishing: 56]  [Article Influence: 2.4]  [Reference Citation Analysis (0)]
53.  Elbashir SM, Lendeckel W, Tuschl T. RNA interference is mediated by 21- and 22-nucleotide RNAs. Genes Dev. 2001;15:188-200.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2287]  [Cited by in F6Publishing: 2246]  [Article Influence: 97.7]  [Reference Citation Analysis (0)]
54.  Fire A, Xu S, Montgomery MK, Kostas SA, Driver SE, Mello CC. Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature. 1998;391:806-811.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 10522]  [Cited by in F6Publishing: 9712]  [Article Influence: 373.5]  [Reference Citation Analysis (1)]
55.  Bernstein E, Caudy AA, Hammond SM, Hannon GJ. Role for a bidentate ribonuclease in the initiation step of RNA interference. Nature. 2001;409:363-366.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3476]  [Cited by in F6Publishing: 3191]  [Article Influence: 138.7]  [Reference Citation Analysis (0)]
56.  Paddison PJ, Caudy AA, Bernstein E, Hannon GJ, Conklin DS. Short hairpin RNAs (shRNAs) induce sequence-specific silencing in mammalian cells. Genes Dev. 2002;16:948-958.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1123]  [Cited by in F6Publishing: 1097]  [Article Influence: 49.9]  [Reference Citation Analysis (0)]
57.  Filleur S, Courtin A, Ait-Si-Ali S, Guglielmi J, Merle C, Harel-Bellan A, Clézardin P, Cabon F. SiRNA-mediated inhibition of vascular endothelial growth factor severely limits tumor resistance to antiangiogenic thrombospondin-1 and slows tumor vascularization and growth. Cancer Res. 2003;63:3919-3922.  [PubMed]  [DOI]  [Cited in This Article: ]
58.  Reich SJ, Fosnot J, Kuroki A, Tang W, Yang X, Maguire AM, Bennett J, Tolentino MJ. Small interfering RNA (siRNA) targeting VEGF effectively inhibits ocular neovascularization in a mouse model. Mol Vis. 2003;9:210-216.  [PubMed]  [DOI]  [Cited in This Article: ]
59.  Zhang L, Yang N, Mohamed-Hadley A, Rubin SC, Coukos G. Vector-based RNAi, a novel tool for isoform-specific knock-down of VEGF and anti-angiogenesis gene therapy of cancer. Biochem Biophys Res Commun. 2003;303:1169-1178.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 82]  [Cited by in F6Publishing: 85]  [Article Influence: 4.0]  [Reference Citation Analysis (0)]
60.  Detwiller KY, Fernando NT, Segal NH, Ryeom SW, D'Amore PA, Yoon SS. Analysis of hypoxia-related gene expression in sarcomas and effect of hypoxia on RNA interference of vascular endothelial cell growth factor A. Cancer Res. 2005;65:5881-5889.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 106]  [Cited by in F6Publishing: 114]  [Article Influence: 6.0]  [Reference Citation Analysis (0)]
61.  Kwon HS, Shin HC, Kim JS. Suppression of vascular endothelial growth factor expression at the transcriptional and post-transcriptional levels. Nucleic Acids Res. 2005;33:e74.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 28]  [Cited by in F6Publishing: 28]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
62.  Senn C, Hangartner C, Moes S, Guerini D, Hofbauer KG. Central administration of small interfering RNAs in rats: a comparison with antisense oligonucleotides. Eur J Pharmacol. 2005;522:30-37.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 36]  [Cited by in F6Publishing: 31]  [Article Influence: 1.6]  [Reference Citation Analysis (0)]
63.  Read ML, Logan A, Seymour LW. Barriers to Gene Delivery Using Synthetic Vectors. Adv Genet. 2005;53PA:19-46.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 47]  [Cited by in F6Publishing: 51]  [Article Influence: 2.7]  [Reference Citation Analysis (0)]
64.  Amarzguioui M, Rossi JJ, Kim D. Approaches for chemically synthesized siRNA and vector-mediated RNAi. FEBS Lett. 2005;579:5974-5981.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 139]  [Cited by in F6Publishing: 113]  [Article Influence: 5.9]  [Reference Citation Analysis (0)]