• 沒有找到結果。

I. Introduction and literature review

1.3 VEGF and cancers

1.3.1 Vascular endothelial growth factor family, VEGF family

VEGF-A (commonly referred to as VEGF) was first identified by Dvorak et al. (2002) as a

vascular permeabilitity-inducing factor secreted by tumor cells, and thus referred to as vascular

permeability factor (VPF). Ferrara et al. (1996) later isolated and cloned VEGF-A as an endothelial

specific mitogen. VEGF is a major inducer of angiogenesis and it is structurally related to PlGF

(placenta growth factor), VEGF-B, VEGF-C, VEGF-D and Orf virus derived VEGF (also called

VEGF-E) (Ferrara and Davis-Smyth, 1997; Erikson and Alitalo, 1999; Persico et al., 1999; Achen

et al., 1998; Ogawa et al., 1998; Meyer et al., 1999). The biological functions of the VEGF family

29

are mediated by activation of three structurally homologous tyrosine kinase receptors, VEGFR-1,

VEGFR-2, and VEGFR-3 (Neufeld et al., 1999; Veikkola et al., 2000; Aprelikova et al., 1992)

1.3.1.1 VEGF-A

VEGF-A is a 45-kDa homodimeric glycoprotein with a diverse range of angiogenic activities.

The VEGF-A gene undergoes alternative splicing to yield mature isoforms of 121, 165, 189, and

206 amino acids (Houck, 1991; Tischer et al., 1991). In addition, some less commonly expressed

variants have also been identified (VEGF145 and VEGF183) (Neufeld et al., 1999).

VEGF121 is freely secreted, whereas the largest isoforms (VEGF189 and VEGF206) are

sequestered in the extracellular matrix (ECM) and require cleavage by proteases for their activation

(Dvorak, 2002). VEGF165 exists in both a soluble and an ECM-bound form (Keyt et al., 1996). The

ECM-bound isoforms of VEGF-A, VEGF-C, and VEGF-D can be released in a diffusible form by

plasmin cleavage at the C-terminus, which generates a bioactive fragment (Park et al., 1993;

McColl et al., 2003). Alternatively, VEGF can be released from the ECM by MMP-9 to initiate the

angiogenic switch (Bergers et al., 2000). VEGF165 is the predominant isoform and is commonly

overexpressed in a variety of human solid tumors. In mice, homozygous or heterozygous deletion of

the VEGF gene is embryonically lethal, resulting in defects in vasculogenesis and cardiovascular

abnormalities, demonstrating that VEGF is essential for development (Carmeliet et al., 1996;

Ferrara, 1996). VEGF-A is also important to a number of postnatal angiogenic processes, including

30

wound healing, ovulation, menstruation, maintenance of blood pressure, and pregnancy (Brown,

1992). VEGF-A has also been linked to several pathologic conditions associated with increased

angiogenesis, including arthritis, psoriasis, macular degeneration, and diabetic retinopathy (Ferrara

et al., 2003). In healthy conditions, endothelial cells survive for prolonged periods owing to the low

level of VEGFA expression by the vasculature, which acts as a survival signal. Hence, when

quiescent endothelial cells are deprived of this trophic VEGFA signal, they become dysfunctional

(pro-thrombotic), lose their vasorelaxing activity (nitric oxide release) or even disappear altogether

(which leads to bleeding) (Baffert et al., 2006; Gerber et al., 1999; Lee et al., 2007). Furthermore,

large amounts of VEGFA are produced in healthy conditions, which suggest that VEGFA is needed

for endothelial-cell homeostasis (Rudge et al., 2007). VEGFA binds to FLT1 with an affinity

(dissociation constant; Kd ~2–10 pM) that is much higher than for FLK1 (Sawano et al., 2001). Yet,

VEGFA induces weaker tyrosine-kinase activity in FLT1, possibly because of an inhibitory

sequence in the juxtamembrane domain that represses FLT1 activity (Seetharam et al., 1995;

Waltenberger et al., 1994; Gille et al., 2000). This weak tyrosine-kinase activity of FLT1 and its

high affinity for VEGFA have led to the development of a model in which FLT1 acts as a decoy

receptor and modulates angiogenesis through its ability to sequester VEGFA, and thereby reduces

signalling through FLK1 (Park et al., 1994). Additional functional diversity may occur through the

formation of dimers between VEGF family members. For example, VEGF-A may form

heterodimers with either PlGF (Cao et al., 1996) or VEGF-B (Olofsson et al., 1996). In humans,

31

because of the alternative splice variants, this has the potential to create enormous diversity and

presents a challenge to assessing the functional consequences of heterodimerization (Birkenhager et

al., 1996). For example, the six human VEGF-A and four human PlGF isoforms could theoretically

form 24 VEGF-PlGF heterodimers, and VEGF-A could theoretically form 12 different

combinations with the two human VEGF-B variants.

1.3.1.2. VEGF-B and PlGF

VEGF-B and PlGF null mice display no defects in embryonic vasculogenesis or developmental

abnormalities, suggesting that the role of PlGF and VEGF-B may be redundant (Carmeliet, 2001).

However, loss of PlGF impairs angiogenesis, plasma extravasation, and collateral growth during

ischemia, inflammation, wound healing, and tumor growth, suggesting a role for PlGF in pathologic

states in the adult. VEGF-B exists as two isoforms (VEGF167 and VEGF 186) that bind to FLT1

and to neuropilin 1 (Neufeld, 2001). Neuropilin 1 is widely expressed in various tissues (Nash, 2006)

including endothelial and mural cells, brown fat, skin, lung, placenta, brain and retina, but it is

particularly abundant in the heart and skeletal muscle (Li, 2001). Despite some evidence for

VEGFB activity in vitro, genetic studies have revealed that VEGFB-deficient mice are healthy and

fertile, and do not display vascular defects, which indicates that VEGFB is redundant in

angiogenesis in the developing embryo and healthy adult in vivo (Li et al., 2008; Aase et al., 2001;

Bellomo et al., 2000). The role of VEGFB in tumor growth remains largely elusive. VEGFB has

32

been detected in a wide range of tumors, including meningioma, and colorectal, lung and breast

cancer (Andre et al., 2000; Donnini et al., 1999; Niki et al., 2000). Moreover, VEGFB expression

correlates with microvascular density in oral squamous cell carcinoma (Shintani et al., 2004).

1.3.1.3 VEGF-C and VEGF-D

The VEGF homologs VEGF-C and VEGF-D play key roles during embryonic and postnatal

lymphangiogenesis. Homozygous deletion of the VEGF-C gene in mice is embryonically lethal and

heterozygous deletion results in postnatal defects associated with defective lymphatic development

(Karkkainen et al., 2004). Both factors induce lymphangiogenesis in transgenic mouse models

(Jeltsch et al., 1997; Veikkola et al., 2001). VEGF-C and VEGF-D may also play a role in new

blood vessel growth as well, especially during pathological states such as tumor growth.

1.3.1.4. VEGF-E

VEGF-E is not a mammalian VEGF homolog, but rather a viral protein encoded by the

parapoxvirus Orf virus, which preferentially utilizes kinase-insert domain-receptor (KDR)/fetal

liver kinase-1 (Flk-1) receptor and carries a potent mitotic activity without heparin-binding domain

(Ogawa et al., 1998). VEGF-E shares 22% sequence identity to VEGF-A.

1.3.2 VEGF receptors

33

VEGF ligands mediate their angiogenic effects via several different receptors. Two receptors

were originally identified on endothelial cells and characterized as the specific tyrosine kinase

receptors VEGFR-1 (also referred to as fms-like tyrosine kinase 1 [Flt-1]) (Shibuya et al., 1990).

VEGFR-2 (also referred to KDR, (Terman, 1992) and the murine homologue, Flk-1) (Matthews et

al. 1991). These two receptors share 44% homology and possess a characteristic structure consisting

of seven extracellular immunoglobulin-like domains, a single transmembrane domain, and a

consensus tyrosine kinase domain interrupted by a kinase insert domain (Shibuya et al., 1990;

Terman, 1991). More recently, an additional tyrosine kinase receptor, VEGFR-3 (also referred to as

fms-like tyrosine kinase 4 [Flt-4]), was identified and has been found to be primarily associated

with lymphangiogenesis (Kaipainen et al., 1995; Paavonen et al., 2000). The various members of

the VEGF family have differing binding specificities for each of these receptors. All of the

VEGF-A isoforms bind to both VEGFR-1 and VEGFR-2, whereas PlGF-1, PlGF-2 and VEGF-B

are specific for VEGFR-1 binding and activation (Park et al., 1994; Olofsson et al., 1998; Silvestre

et al., 2003). Naturally occurring heterodimers of VEGF-A and PlGF have also been identified that

can bind to and activate VEGFR-2 (Cao et al., 1996; DiSalvo et al., 1995). VEGF-E specifically

interacts with VEGFR-2 whereas VEGF-C and VEGF-D interact with both VEGFR-3 and

VEGFR-2 (Shibuya et al., 2003; Achen et al., 1998; Joukov et al., 1996).

1.3.2.1 VEGF-R1

34

VEGFR-1 is a receptor for VEGF-A and has the unique ability to bind VEGF-B and PlGF.

VEGFR-1 is critical for physiologic and developmental angiogenesis. VEGFR-1 null mice die in

utero between days 8.5 to 9.5 due to excessive hemangioblast proliferation and poor organization of

vascular structures (Fong et al., 1995). VEGFR-1 was initially thought to be a negative regulator of

VEGF activity either by acting as a decoy receptor for VEGF (Hiratsuka et al., 1998) or by

downregulating VEGFR-2–mediated signaling (Dunk et al., 2001). VEGF-mediated stimulation of

VEGFR-1 autophosphorylation and signaling in endothelial cells is weak when compared to

signaling through VEGFR-2 (Waltenberger et al., 1994). A repressor motif has been identified in

the juxtamembrane region of VEGFR-1 that impairs phosphatidylinositol 3-kinase (PI3K) signaling

and endothelial cell migration in response to VEGF stimulation (Gille et al., 2000; Zeng et al.,

2001). However, other studies have indicated that VEGFR-1 has a positive, functional role in

certain cell types, participating in monocyte migration (Barleon et al., 1996; Clauss et al., 1996),

recruitment of endothelial cell progenitors (Lyden et al., 2001) increasing the adhesive properties of

natural killer cells (Chen et al., 2002) and inducing growth factors from liver sinusoidal endothelial

cells (LeCouter et al., 2003). A recent study (Autiero et al, 2003) showed that activation of

VEGFR-1 by PlGF resulted in transphosphorylation of VEGFR-2 in endothelial cells co-expressing

these receptors. Furthermore, VEGF/PlGF heterodimers are capable of activating intramolecular

VEGFR cross talk through formation of VEGFR-1/VEGFR-2 heterodimers. Other recent studies

have shown that during pathologic conditions such as tumorigenesis, VEGFR-1 is a potent, positive

35

regulator of angiogenesis (Hiratsuka et al., 2001). A naturally occurring, alternatively spliced

soluble form of VEGFR-1 (sVEGFR-1) also exists (Kendall et al., 1993). sVEGFR-1 may function

to reduce or modulate endogenous VEGF or PlGF activity.

1.3.2.2. VEGF-R2

VEGFR-2 mediates the majority of the downstream effects of VEGF-A in angiogenesis,

including microvascular permeability, endothelial cell proliferation, invasion, migration, and

survival (Zeng et al., 2001; Millauer et al., 1993). The importance of VEGFR-2 in vasculogenesis is

demonstrated by the fact that hetero and homozygous knockout mice die in utero of defects in blood

island formation and vascular development (Shalaby et al., 1995). VEGFR-2–mediated proliferation

of endothelial cells involves activation of a phospholipase C gamma-protein kinase C-Raf-MAP

kinase signaling pathway, whereas survival and migration is believed to involve PI3K and focal

adhesion kinase, respectively (Veikkola et al., 2000; Abedi et al., 1997).

1.3.2.3 VEGFR3

VEGFR-3 is a receptor tyrosine kinase originally cloned from a human leukemia cell line and

human placenta (Pajusola et al., 1993; Galland et al., 1993). VEGFR-3 preferentially binds

VEGF-C and VEGFD. VEGFR-3 is expressed throughout the embryonic vasculature, but during

development and in the adult, its expression is limited to lymphatic endothelial cells (Kaipainen et

al., 1995). Homozygous deletion of the VEGFR-3 gene in mice leads to embryonic death at day 10

36

to 12.5, with an underdeveloped yolk sac, poor perineural vasculature, and pericardial fluid

accumulation (Dumont et al., 1998). VEGFR-3 activation and upregulation of its ligands have been

observed in certain neoplastic conditions, including breast cancer and melanoma with elevated

levels of VEGF-C or VEGF-D associated with lymph node metastasis in patients (Achen et al.,

2001; Valtola et al., 1999; Pepper et al., 2003).

1.3.3 Functions of VEGF

1.3.3.1 Vessels Permeability

VEGF is also termed vascular permeability factor (VPF) (Dvorak et al., 1995). In fact, VEGF

is one of the most potent inducers of vascular permeability known - 50,000-fold more potent than

histamine (Dvorak, 2002). This ability to enhance microvascular permeability is one of the most

important properties of VEGF, especially with regards to the hyperpermeability of tumor vessels

that is thought to be largely attributable to tumor cell expression of VEGF. It has been suggested

that the increase in permeability results in the leakage of several plasma proteins, including

fibrinogen and other clotting proteins. This can lead to the deposition of fibrin in the extravascular

space, which subsequently retards the clearance of edema fluid and transforms the normally

antiangiogenic stroma of normal tissues into a proangiogenic environment (Dvorak et al., 1995;

2002). VEGF increases permeability in a variety of vascular beds, including the skin, peritoneal

wall, mesentery, and diaphragm, and can lead to pathologic conditions such as malignant ascites

37

and malignant pleural effusions (Yoshiji et al., 2001; Yuan et al., 1996).

1.3.3.2 Endothelial cell activation

VEGF exerts a number of different effects which include changes in endothelial cell

morphology, cytoskeleton alterations, and stimulation of endothelial cell migration and growth.

VEGF causes increased expression of a variety of different endothelial cell genes, including

procoagulant tissue factor and fibrinolytic pathway proteins, such as urokinase, tissue-type

plasminogen activator, type 1 plasminogen activator inhibitor, and urokinase inhibitor; and matrix

metalloproteases; GLUT-1 glucose transporter; nitric oxide synthase; and integrins (Dvorak et al.,

2002; Zachary, 2001; Eliceiri et al., 2001; 1999; Brooks et al., 1994).

1.3.3.3 Endothelial cell survival

VEGF was first shown to act as a survival factor for retinal endothelial cells (Alon et al., 1995).

In vitro, VEGF has been shown to inhibit apoptosis by activating the PI3K-Akt pathway (Gerber et

al., 1998; Gerber et al., 1998; Thakker et al., 1999) in addition to upregulating antiapoptotic

proteins such as bcl-2 and A1. VEGF has also been shown to activate focal adhesion kinase (FAK)

and associated proteins that have been shown to maintain survival signals in endothelial cells

(Abedi et al., 1997; Zachary et al., 2001).

1.3.3.4 Endothelial cell proliferation

38

VEGF is a mitogen for endothelial cells. This endothelial cell proliferation appears to involve

VEGFR-2–mediated activation of extracellular kinases Erk1/2 in addition to another member of the

MAP kinase family, JNK/SAPK (Zachary et al., 2001; Meadows et al., 2001).

1.3.3.5 Endothelial cell invasion and migration

Degradation of the basement membrane is necessary for endothelial cell migration and

invasion and is an important early step in the initiation of angiogenesis. VEGF induces a variety of

enzymes and proteins important in the degradation process, including matrix-degrading

metalloproteinases, metalloproteinase interstitial collagenase, and serine proteases such as

urokinase-type plasminogen activator (uPA) and tissue-type plasminogen activator (TTPA)

(Zachary et al., 2001; Choong et al., 2003). Activation of these various compounds leads to a

prodegradative environment that facilitates migration and sprouting of endothelial cells (Ferrara et

al., 1997). The intracellular mechanisms by which VEGF leads to increased endothelial cell

migration are not entirely clear, but appear to involve FAK-associated signaling leading to focal

adhesion turnover and actin filament organization as well as p38 MAPK-induced actin

reorganization (Abedi et al., 1997; Zachary et al., 2001)

1.3.4 Expression of VEGF on tumor cells

Several studies have reported the presence of VEGFRs on liquid and solid tumor cells,

39

including those of non–small cell lung carcinoma, melanoma, prostate carcinoma, leukemia,

mesothelioma, and breast carcinoma (Dias et al., 2001; Bellamy et al., 1999; Decaussin et al., 1999;

Dias et al., 2000; Ferrer et al., 1999; Hayashibara et al., 2001; Lacal et al., 2000; Price et al., 2001).

Various VEGF ligands support tumor growth, not only by inducing angiogenesis, but also by acting

directly through VEGFRs expressed by tumor cells. Moreover, since the vast majority of solid

tumors and a variety of hematologic malignancies have the capacity to express VEGF, expression

of VEGFRs by tumor cells implicates a potential role for VEGF/VEGFR autocrine loops in these

tumors.

1.3.5 Expression of VEGF on OSCCs

VEGF expression is up-regulated significantly during the transition from NOM, through

dysplasia to OSCC. For the dysplastic lesions, no correlation is found between VEGF expression

and grade of dysplasia (Johnstone et al., 2007). The head and neck squamous cell carcinoma

(HNSCC) tumor cells express VEGFR-1, VEGFR-2, and VEGFR-3 in all specimens evaluated.

Staining for all 3 receptors is also found on tumor associated macrophages and fibrobasts, except

that VEGFR-2 is not present on fibroblasts. Staining intensity for VEGFR-1 and VEGFR-2 is

significantly higher in tumor cells and macrophages than in vascular endothelial cells (VECs)

stained for the same receptor. In HNSCC, as well as other tumor systems, VEGF has been shown to

be expressed by tumor cells and to induce proliferation of adjacent VECs via a paracrine

40

mechanism (Benefield et al., 1996; Petruzzelli et al., 1997). Recent evidence has demonstrated that

VEGF may also have a direct effect on tumor cell activity (Masood et al., 2001). These data suggest

a possible autocrine mechanism for VEGF-regulated vasculogenesis and tumorigenesis.

相關文件