Fibroblast Growth Factor-2 Antagonist and ...

3 downloads 0 Views 506KB Size Report
includes also guinea pig apexin, rat, human, and murine neuronal pentraxins 1 (NP1 or NPTX1) and NP2 (also called. Narp or NPTX2), and the putative integral ...
Current Pharmaceutical Design, 2009, 15, 3577-3589

3577

Fibroblast Growth Factor-2 Antagonist and Antiangiogenic Activity of Long-Pentraxin 3-Derived Synthetic Peptides D. Leali, P. Alessi, D. Coltrini#, M. Rusnati, L. Zetta§ and M. Presta* Unit of General Pathology and Immunology and #Unit of Histology, Department of Biomedical Sciences and Biotechnology, School of Medicine, University of Brescia, Italy; §Istituto Per lo Studio delle Macromolecole, Consiglio Nazionale delle Ricerche, 20133 Milan, Italy Abstract: Angiogenesis and inflammation are closely integrated processes. Fibroblast growth factor-2 (FGF2) is a prototypic angiogenesis inducer belonging to the family of the heparin-binding FGF growth factors. FGF2 exerts its proangiogenic activity by interacting with various endothelial cell surface receptors, including tyrosine kinase receptors, heparan-sulfate proteoglycans, and integrins. A tight cross-talk exists between FGF2 and the inflammatory response in the modulation of blood vessel growth. Pentraxins act as soluble pattern recognition receptors with a wide range of functions in various pathophysiological conditions. The long-pentraxin PTX3 shares the C-terminal pentraxin-domain with shortpentraxins and possesses a unique N-terminal domain. These structural features indicate that PTX3 may have distinct biological/ligand recognition properties when compared to short-pentraxins. Co-expression of PTX3 and FGF2 has been observed in different inflammation/angiogenesis-dependent diseases. PTX3 binds FGF2 with high affinity and specificity. The interaction prevents the binding of FGF2 to its cognate tyrosine kinase receptors, leading to inhibition of the angiogenic activity of the growth factor. This suggests that PTX3 may exert a modulatory function by limiting the angiogenic activity of FGF2. An integrated approach that utilized PTX3 fragments, monoclonal antibodies, and surface plasmon resonance analysis has identified the FGF2-binding domain in the unique N-terminal extension of PTX3. On this basis, PTX3-derived synthetic peptides have been designed endowed with a significant antiangiogenic activity in vitro and in vivo. They may provide the basis for the development of novel antiangiogenic FGF2 antagonists.

Keywords: Angiogenesis, FGF2, innate immunity, pentraxin, tumor, synthetic peptides, NMR, modeling. 1. FGF2: ANGIOGENESIS AND INFLAMMATION 1.1. FGF2 as an Angiogenic Growth Factor Angiogenesis, the process of new blood vessel formation from pre-existing ones, plays a key role in various physiological and pathological conditions, including embryonic development, wound repair, inflammation, and tumor growth [1]. Angiogenesis is a multi-step process that begins with the degradation of the basement membrane by activated endothelial cells (ECs) that will migrate and proliferate, leading to the formation of solid EC sprouts into the stromal space. Then, vascular loops are formed and capillary tubes develop with formation of tight junctions and deposition of a new basement membrane [2]. The local, uncontrolled release of angiogenic growth factors and/or alterations of the production of natural angiogenic inhibitors, with a consequent alteration of the angiogenic balance [3], are responsible for the uncontrolled EC proliferation that takes place during tumor neovascularization and in angiogenesis-dependent diseases [4]. Numerous inducers of angiogenesis have been identified, including the members of the vascular endothelial growth factor (VEGF) family, angiopoietins, transforming growth *Address correspondence to this author at the General Pathology and Immunology, Department of Biomedical Sciences and Biotechnology, Viale Europa 11, 25123 Brescia, Italy; Tel: ++39-30-3717311; Fax: ++39-30-3701157; E-mail: [email protected] 1381-6128/09 $55.00+.00

factor- and - (TGF- and -), platelet-derived growth factor (PDGF), tumor necrosis factor- (TNF-), interleukins (ILs), chemokines, and members of the fibroblast growth factor (FGF) family. Twenty-three structurally-related members of the FGF family have been identified [5]. FGFs are pleiotropic factors acting on different cell types, including ECs, following interaction with heparan-sulfate proteoglycans (HSPGs) and tyrosine kinase FGF receptors (FGFRs). FGFRs belong to the subclass IV of membrane-spanning receptors, are encoded by four distinct genes, and their structural variability is increased by alternative splicing [6]. FGFR1 [7], and less frequently FGFR2 [8] are expressed by ECs, whereas the expression of FGFR3 or FGFR4 has never been reported in endothelium. Among the FGF family members, FGF2 represents the prototypic and best characterized proangiogenic factor. FGF2 expression is augmented at sites of chronic inflammation [9-11], after tissue injury [12], and in different types of human cancer [13]. In vitro, FGF2 binds all FGFRs, with preferential activation of the alternative spliced IIIc form in FGFRs 1-3 [14]. FGF2/FGFR interaction causes receptor dimerization and autophosphorylation of specific tyrosine residues located in the intra-cytoplasmic tail of the receptor. This in turn leads to complex signal transduction pathways and activation of a “pro-angiogenic phenotype” in ECs (reviewed in [13]). In vivo, FGF2 has been shown to

© 2009 Bentham Science Publishers Ltd.

3578 Current Pharmaceutical Design, 2009, Vol. 15, No. 30

induce neovascularization in a variety of animal models, including the chick embryo chorioallantoic membrane (CAM) assay, the rodent cornea assay, the subcutaneous (s.c.) Matrigel plug assay in mice, and the zebrafish yolk membrane (ZFYM) assay [13, 15]. FGF2 can exert its effects on ECs via a paracrine mode consequent to its release by tumor, stromal, and inflammatory cells and/or by mobilization from the ECM. On the other hand, endogenous FGF2 produced by ECs may play important autocrine, intracrine, or paracrine roles in angiogenesis and in the pathogenesis of vascular lesions, including Kaposi’s sarcoma and hemangiomas (see [16] and references therein). 1.2. FGF2-Dependent Angiogenesis and Inflammation Inflammation is the response of a vascularized tissue to sub-lethal injury, designed to destroy or inactivate invading pathogens, remove waste and debris, and permit restoration of normal function, either through resolution or repair. Angiogenesis and inflammation are closely integrated processes in a number of physiological and pathological conditions, including wound healing, psoriasis, diabetic retinopathy, rheumatoid arthritis, atherosclerosis, and cancer [1, 17, 18]. Inflammation may promote FGF2-dependent angiogenesis. Inflammatory cells, including mononuclear phagocytes [19, 20], CD4+ and CD8+ T lymphocytes [21, 22], and mast cells [23] can express FGF2. Moreover, osmotic shock and shear stress induce the release of FGF2 from ECs [24, 25]. FGF2 production and release from ECs are also triggered by interferon (IFN)- plus IL-2 [26], IL-1 [27], and nitric oxide (NO) [28]. Indeed, the pro-angiogenic effects exerted by NO and NO-inducing molecules are due, at least in part, to the NO-mediated FGF2 upregulation in ECs [29]. Similarly, prostaglandin E2-induced angiogenesis is mediated by the activation of EC-surface FGFR1 following mobilization of FGF2 sequestered by the ECM [30]. Thus, inflammatory mediators can activate the endothelium to synthesize and release FGF2 that, in turn, will stimulate angiogenesis by an autocrine mechanism of action. The inflammatory response may also cause cell damage, fluid and plasma protein exudation, and hypoxia. EC damage results in increased FGF2 production and release [31]; exudated fibrin(ogen) can bind FGF2 and enhances its biological activity [32, 33]; hypoxia upregulates the production of FGF2 in mononuclear phagocytes [19] and in vascular pericytes [34]. Furthermore, hypoxia increases EC responsiveness to FGF2 by promoting HSPG synthesis [35]. Conversely, by interacting with ECs, FGF2 may amplify both the inflammatory and the angiogenic responses by inducing vasoactive effects [36], vascular permeability [37], and the recruitment of an inflammatory infiltrate [38]. Moreover, FGF2-stimulated ECs upregulate the synthesis of various monocyte chemoattractants, including VEGF [39], the chemokine CCL2 [40, 41], and osteopontin (OPN) [42]. Relevant to this point, OPN induces neovascularization [42] by promoting the release of the pro-angiogenic cytokine IL1 from recruited monocytes/macrophages [43].

Leali et al.

In keeping with the existence of a tight cross-talk between inflammatory and angiogenic responses during FGF2-driven neovascularization, gene expression profiling of FGF2-stimulated murine microvascular ECs has revealed a pro-inflammatory signature characterized by the upregulation of pro-inflammatory cytokine/chemokines and their receptors, EC adhesion molecules, and members of the eicosanoid pathway [44]. Accordingly, we have observed that the early recruitment of mononuclear phagocytes precedes blood vessel formation in FGF2-driven angiogenesis in the s.c. Matrigel plug assay and that monocytes/macrophages play a functional, non-redundant early role in FGF2-driven angiogenesis [44]. It must be pointed out that, together with a proinflammatory signature, FGF2 upregulates also the expression of a variety of angiogenic growth factors in ECs, including FGF2 itself and VEGF [44]. This suggests that FGF2 is able to activate an autocrine loop of amplification of the angiogenic response that, together with the paracrine activity exerted by endothelium-derived cytokines/chemokines on inflammatory cells, will contribute to the modulation of the neovascularization process triggered by the growth factor. 1.3. FGF2 as a Target for Anti-Angiogenic Strategies A significant effort has been directed towards the development of anti-angiogenic agents that prevent the growth of new blood vessels, the monoclonal anti-VEGF antibody bevacizumab and the tyrosine kinase inhibitors sunitinib and sorafenib representing the first FDA approved anti-angiogenic drugs [45]. On the other hand, drug resistance to VEGF blockade may occur following reactivation of angiogenesis triggered by compensatory upregulation of the FGF2/FGFR system in experimental tumor models [46] and in cancer patients [47]. Thus, given its potent angiogenic activity, FGF2 may represent the target for the development of novel anti-angiogenic strategies. The various approaches based on the inhibition of FGF2 have been reviewed extensively elsewhere (see [48] and references therein). Briefly, the angiogenic activity of FGF2 can be neutralized by: i) inhibition of FGF2 production/ release; ii) inhibition of the expression of the various FGF2 receptors in ECs; iii) engagement of the various FGF2 receptors by selected antagonists; iv) sequestration of FGF2 in the extracellular environment; v) interruption of the signal transduction pathways triggered by FGF2 in ECs. Also, as stated above, FGF2 induces a complex “proinflammatory phenotype” in ECs whose blockage may result in the inhibition of FGF2-dependent angiogenesis [48]. Indeed, we have observed that FGF2-mediated angiogenesis is significantly reduced in the CAM assay by both steroidal (hydrocortisone) and non-steroidal (ketoprofen) anti-inflammatory drugs, further implicating inflammatory cells/ mediators in FGF2-dependent neovascularization [44]. Moreover, FGF2-induced neovascularization is inhibited also by M3 protein [44], a murine gammaherpesvirus 68 protein that binds with high affinity to human and mouse CC, CXC and CX3C chemokines and inhibits their activity [49, 50], with potential therapeutic implications in inflammatory conditions [51].

Antiangiogenic Activity of PTX3

Current Pharmaceutical Design, 2009, Vol. 15, No. 30

Interestingly, several ECM and serum components and/or their degradation products affect FGF-driven angiogenesis, including thrombospondin-1 (TSP-1) [52, 53], the fibronectin fragment fibstatin [54], the plasma protein 2macroglobulin [55, 56], PDGF-BB [57], and the chemokines CXCL4 [58] and CXCL13 [59]. In the search for effective FGF2 antagonists, numerous peptides derived from these natural FGF2-binders, FGFRs, and FGF2 itself have been demonstrated to exert an inhibitory activity on the FGF2/FGFR system (Table 1). 2. LONG-PENTRAXIN 3 2.1. PTX3 as a Member of the Pentraxin Family Pentraxins are a superfamily of evolutionarily conserved proteins originally named for their structural organization characterized by a radial pentameric structure. Pentraxins are Table 1.

3579

divided into two subfamilies (short-pentraxins and longpentraxins) sharing a C-terminal pentraxin domain that contains the HxCxS/TWxS pentraxin signature (where x is any amino acid) [60]. Long-pentraxins differ from shortpentraxins for the presence of an unrelated N-terminal domain coupled to the C-terminal domain [61] (Fig. 1A). C-reactive protein (CRP) and serum amyloid P component (SAP) constitute the short-pentraxin arm of the superfamily. These “classical” pentraxins are acute phase proteins in man and mouse, respectively, and are produced in liver in response to inflammatory mediators, most prominently IL-6 [62, 63]. They are involved in the innate resistance to microbes and scavenging of cellular debris and ECM components [60, 64-66]. Pentraxin 3 (PTX3, also called TSG-14) is the prototypic member of the long-pentraxin subfamily [67, 68] that

Synthetic Peptides Endowed with FGF2 Antagonist Activity

Protein of Origin

Peptide(s)

Target

References

FGF1

FGF1(112-147) and related peptides; FGF1(99-108); FGF1 mimetics (6 peptides studied)

FGFR

[117-119]

FGF2(48-58) (FREG)

FGF2

[120]

FGF2(38-61); FGF2(82-101); FGF2-derived DGR-containing peptides (4 peptides studied)

ND

[121]

FGF2(119-126)

FGF2

[122]

FGF2(68-77)

FGFR

[123]

FGF2(24-68) (Peptide D); FGF2(93-120) (Peptide N); FGF2(106-115)

FGFR

[124, 125]

FGF2(103-146)

FGFR

[126]

F2A4-K-NS

FGFR

[127]

16-24 mer peptides based on FGF2 molecular modeling

FGFR

[128, 129]

FGF5

FGF5(95-104) (peptide P3)

FGFR

[130]

FGFs (10-11 loop)

dekafins (homologous to the NCAM FGFR-binding region)

FGFR

[131]

Myelin Basic Protein

MBP(152-167)

FGFR

[132]

FGL

FGFR

[133-136]

FRM-10 and cyclic FRM-10. FRM-13

FGFR

[134, 136, 137]

DekaCAM

FGFR

[131, 134, 136]

BCL

FGFR

[134, 136, 138]

Encamin A, C, E

FGFR

[134, 139]

N-cadherin

Extracellular domain-4 mimetics (2 peptides studied)

FGFR

[119]

PF4 (CXCL4)

PF4(47-70)*

FGF2

[115, 140]

PTX3 N-terminus

PTX3(82-110) (and related peptides, including ARPCA*; see text for further details)

FGF2

[73, 111]

Epitope sequence, FGF2(13-18), FGF2(119-126), FGF2(120-125)

FGFR

[141]

Peptide P7*

FGF2

[116]

C19 (3 peptides studied)

FGFR

[142-144]

Peptide P2

FGFR

[145]

4N1K

ND

[146]

(type III repeats-derived peptides) (6 peptides studied)

FGF2

[147]

FGF2

NCAM(681-695)

Random phage epitope library

TSP-1 *peptide establishing hydrophobic interactions with FGF2. ND: mechanism of action not defined.

3580 Current Pharmaceutical Design, 2009, Vol. 15, No. 30

includes also guinea pig apexin, rat, human, and murine neuronal pentraxins 1 (NP1 or NPTX1) and NP2 (also called Narp or NPTX2), and the putative integral membrane pentraxin NRP [60] for a review]. Long-pentraxins have been found in Xenopus [61], Drosophila melanogaster [69], and zebrafish (Danio Rerio) [60]. The human PTX3 gene has been identified using differential screening of cDNA libraries constructed from IL1-stimulated human umbilical vein ECs [67] and from normal FS-4 fibroblasts stimulated by TGF- and TNF- [70]. Unlike the CRP and SAP genes that map on chromosome 1, the human PTX3 gene is localized on chromosome 3, band q25. The PTX3 gene comprises three exons, the first exon extending to nucleotide 197, the second one covering nucleotides 198-599, and the third one extending from nucleotide 600 to the 3’-terminus. The first two exons encode for the signal peptide and the N-terminal domain of PTX3 protein whereas the third exon encodes for its Cterminal domain, matching exactly the second exon of the short-pentraxin genes [71]. Thus, PTX3 appears to represent a fusion between regions encoding a specific N-terminal polypeptide of unknown function and a C-terminal polypeptide homologous to short-pentraxins. 2.2. PTX3 Protein Structure The human PTX3 protomer is a 381 amino acid glycoprotein, including a 17 amino acid signal peptide for secretion. The PTX3 primary sequence is highly conserved among animal species (human and murine PTX3 share 92% of conserved amino acid residues), suggesting a strong evolutionary pressure to maintain its structure-function relationships [60]. As the other members of the longpentraxin subfamily, PTX3 is composed of an unique Nterminal domain (spanning amino acid residues 1-178) and of a C-terminal 203 amino acid domain highly homologous among the various members of the pentraxin family (57% of conserved amino acids with short-pentraxins CRP and SAP). The human protein has a unique N-glycosylation site at Asn220 fully occupied by complex type-oligosaccharides, mainly fucosylated and sialylated biantennary sugars [72]. These negatively charged sialic acid residues may establish ionic interactions with polar and basic amino acids located far from the Asn220 N-glycosylation site [72]. No crystallographic data are available for the N-terminal portion of long-pentraxins. Consensus secondary structure prediction of PTX3 N-terminus has identified four -helix regions connected by short loops spanning amino acid residues 55-75 (A), 78-97 (B), 109-135 (C), and 144170 (D) [73]. Moreover, amino acid residues 85-91 in B helix form the structural heptad repeat motif (abcdefg), where a and d are hydrophobic residues and e and g represent charged residues [74]. Also, hydrophobic residues repeated with a period of one each three or six amino acids are present within the primary sequence of C and D helices. Thus, B, C, and D helices of the N-terminal PTX3 domain have propensity to be in a coiled-coil conformation [see [75] for further details]. The high similarity of the primary sequence of the PTX3 C-terminus with short-pentraxins allowed instead to produce

Leali et al.

a predicted structure for the C-terminal PTX3 region (residues 179-381) using the crystallographic structure of CRP as template [75]. The model of PTX3 C-terminus presents a hydrophobic core composed by two anti-parallel -sheets organised in a typical -jelly roll topology [76]. A similar model was described when PTX3 was modelled on the tertiary structure of SAP [63, 77]. The mature PTX3 protein contains nine cysteine residues: three residues are located within the N-terminal region (Cys47, Cys49, and Cys103) whereas six residues are in the C-terminal domain (Cys179, Cys210, Cys271, Cys317, Cys318, and Cys357). Cys210 and Cys271 are highly conserved among pentraxins and, based on the homology with CRP and SAP, are predicted to be engaged in an intrachain disulfide bond [61, 78]. The other Cys residues are involved in intra- and inter-chain disulfide bonds, thus determining the quaternary structure of PTX3 [76]. Indeed, PTX3 shows a complex quaternary structure with subunits assembled into high order oligomers [78]. Recently, the oligomeric assembly of PTX3 has been resolved [76], experimental data demonstrating that human recombinant PTX3 is mainly composed of covalently linked octamers (Fig. 1B). The multimeric organization of PTX3 is essential for the organization of the ECM of the cumulus oophorus, the tetramer being the smallest PTX3 oligomer still retaining full functionality in cumulus matrix assembly and stabilization [76]. 2.3. PTX3 Biological Functions and Interactions PTX3 is produced at the inflammatory site in response to inflammatory cytokines and bacterial components [67, 68, 79, 80]. Accordingly, the proximal 1317-bp promoter of the human PTX3 gene is responsive to TNF- and IL-1, but not to IL-6. Multiple binding sites for various transcription factors have been identified in this sequence, including NFIL-6, AP-1, Pu1, PEA-3, Ets-1, and Sp1 sites. Also, NF-kB sites are essential for induction by IL-1 and TNF- [81, 82]. Various cell types express PTX3, including ECs [67], macrophages [80], microglia [83], dendritic cells [84], adipocytes [85], fibroblasts [68], and myoblasts [77]. Also, PTX3 is upregulated during vasculitis [86], it is present in atherosclerotic plaques [87], and it is expressed by smooth muscle cells (SMCs) isolated from human arterial specimens [88]. Thus, when compared to the liver-produced shortpentraxin CRP, PTX3 may represent a rapid marker for primary local activation of inflammation as well as of innate immunity. Indeed, the levels of circulating PTX3 increase in humans during vasculitis [86], acute myocardial infarction [89], rheumatoid arthritis [90], and systemic inflammatory response syndrome/septic shock [91]. PTX3 is a soluble pattern recognition receptor that may serve as a mechanism of amplification of inflammation and innate immunity with unique non-redundant functions in various physiopathological conditions such as complement activation, protection against opportunistic pathogens, fertility, and angiogenesis [79]. The biological activity of PTX3 is related to its ability to interact with different ligands via its N-terminal or C-terminal domain as a consequence of the modular structure of the protein (see [75] for a review).

Antiangiogenic Activity of PTX3

Current Pharmaceutical Design, 2009, Vol. 15, No. 30

3581

upregulated rapidly in response to proinflammatory cytokines (e.g., IL-1 and TNF-). TSG-6 is comprised of a Link module and a CUB_C domain, both involved in the binding to different ECM components [93, 94]. The coordinated expression of PTX3 and TSG-6 has been described in inflammatory infiltrates and ECs in inflamed tissues [95] and in the cumulus oophorus where they are essential for in vivo fertilization [92, 96, 97]. At present, the possibility that TSG-6 may affect PTX3/FGF2 interaction is under investigation in our laboratory. 3. PTX3/FGF2 INTERACTION 3.1. Biochemical Interaction

Characterization

of

PTX3/FGF2

When assessed for the capacity to interact with a variety of extracellular signaling polypeptides, PTX3 was found to bind FGF2 with high specificity [98]. Under the same experimental conditions, PTX3 did not bind to a wide panel of cytokines, chemokines, and growth factors representative of different classes of soluble polypeptide mediators, showing only a limited interaction with FGF8 (but not with FGF1 or FGF4, all members of the FGF family) and with VEGF. PTX3/FGF2 interaction occurs with high affinity, with a Kd value ranging between 3.0 x 10–7 and 3.0 x 10-8 M depending upon the experimental model adopted [73, 98]. In agreement with the incapacity of short-pentraxins to bind FGF2 [98], an integrated approach that utilized recombinant N-terminal and C-terminal PTX3 fragments, monoclonal antibodies, and surface plasmon resonance analysis identified the FGF2-binding domain in the PTX3 Nterminus [73]. Indeed, the recombinant N-terminal fragment PTX3(1-178) binds FGF2, prevents PTX3/FGF2 interaction, and inhibits the mitogenic activity of FGF2 in ECs. Also, the monoclonal antibody mAb-MNB4, that recognizes the PTX3(87-99) epitope, prevents FGF2/PTX3 interaction and abolishes the FGF2 antagonist activity of PTX3 [73]. The capacity to bind FGF2 and to inhibit its biological activity represents a novel unanticipated function for the N-terminal extension of PTX3. Fig. (1). Schematic representation of PTX3. (A) The pentraxin superfamily: short- and long-pentraxins show a significant homology in their C-terminal pentraxin domain whereas longpentraxins are characterized by unique N-terminal extensions. The 8 amino acid-long pentraxin family signature is highlighted. (B) The PTX3 octamer: interchain disulfide bonds (dotted lines) connecting the N-terminal (white ellipse) or C-terminal (black bar) regions of the eight PTX3 monomers (a-h) are shown. See ref. [76] for details.

Recent observations have shown the ability of PTX3 to bind FGF2, thus acting as a FGF2 antagonist (see below). The coexpression of PTX3 and FGF2 may occur during inflammation, wound healing, atherosclerosis, and neoplasia, thus allowing a fine tuning of the neovascularization process via the production of both angiogenesis inhibitors and stimulators (Fig. 2). Interestingly, PTX3 has been shown to interact also with TSG-6 [92], a 35 kDa-secreted protein

As stated above, FGF2 acts on target cells by interacting with high affinity FGFRs [99] and low affinity HSPGs [100], leading to the formation of HSPG/FGF2/FGFR ternary complexes [101, 102]. PTX3 inhibits the formation of this complex in an experimental model in which the disruption of the complex abolishes FGF2-mediated cell-cell attachment of HSPG-deficient CHO mutants transfected with FGFR1 to wild-type CHO cells expressing HSPGs but not FGFRs [102] (Fig. 3A,B). Furthermore, surface plasmon resonance analysis has shown that PTX3 prevents the binding of FGF2 to the extracellular domain of FGFR1 immobilized to a BIAcore sensorchip (Fig. 3C) but not to immobilized heparin (Fig. 3D), thus suggesting that PTX3 may interact with the FGFR1-binding domain of FGF2. In keeping with these observations, PTX3 prevents the binding of 125I-FGF2 to FGFRs expressed on human SMCs. Also, the interaction of FGF2 with PTX3 or with a soluble form of FGFR1 are mutually exclusive [103]. Thus, as a consequence of PTX3

3582 Current Pharmaceutical Design, 2009, Vol. 15, No. 30

Leali et al.

Fig. (2). PTX3 as an antiangiogenic FGF2 antagonist. Different cell types produce PTX3 (1) and/or cytokines that stimulate PTX3 upregulation in ECs (2). PTX3 may then inhibit the autocrine (3) or paracrine (4) activity exerted by FGF2 on endothelium, thus modulating the angiogenic response.

Fig. (3). Effect of PTX3 on the formation of the HSPG/FGF2/FGFR1 ternary complex. (A) Schematic representation of the FGF2-mediated cell-cell adhesion assay. The interaction of FGFR1-bearing cells with HSPGs of the cell monolayer occurs in the presence (b) but not in the absence (a) of FGF2. FGF2-mediated cell-cell adhesion is prevented by the binding of PTX3 to FGF2 (c). (B) FGFR1-expressing cells were added to a monolayer of HSPG-expressing cells in the presence of FGF2 (30 ng/ml) and increasing concentrations of PTX3. After 2 h at 37°C, FGFR1-expressing cells bound to the monolayer were counted (see [102] for further experimental details). Data are expressed as percentage of FGF2-mediated cell-cell adhesion. (C) Plasmon resonance analysis of the effect of PTX3 on FGF2/FGFR1 and FGF2/heparin interactions. Injection of PTX3 (220 nM) prevents the binding of free FGF2 (50 nM) to the FGFR1-coated Biacore sensorchip (upper panel) but not the heparin-coated sensorchip (lower panel).

Antiangiogenic Activity of PTX3

interaction, the angiogenic activity of FGF2 is inhibited both in vitro and in vivo. 3.2. Biological Consequences of PTX3/FGF2 Interaction PTX3 interaction inhibits the mitogenic activity exerted in vitro by FGF2 in bovine, murine, and human ECs. As stated above, this appears to be due to the capacity of PTX3 to prevent the binding of FGF2 to FGFRs rather than to a direct action of PTX3 on ECs. Indeed, PTX3 does not inhibit EC proliferation triggered by various mitogens (including serum, diacylglycerol, epidermal growth factor, phorbol ester, or VEGF) and PTX3 pretreatment does not affect EC responsiveness to FGF2 [98]. In keeping with the in vitro observations, PTX3 inhibits FGF2-driven neovascularization in different animal models, including the chick embryo CAM assay [98], the ZFYM assay performed on zebrafish embryos [15], and the Matrigel plug assay in mice (M. Presta, unpublished observations). Experimental evidences support the hypothesis that PTX3 may also inhibit FGF2-driven tumor angiogenesis. FGF2-overexpressing mouse aortic endothelial FGF2-TMAE cells are characterized by the capacity to generate opportunistic vascular lesions in nude mice [104]. When FGF2-T-MAE cells were stably transfected with an expression vector harboring the full length human PTX3 cDNA, these lesions showed a reduced rate of growth when compared to tumors originated by parental cells [98]. Similarly, PTX3 protein caused a significant inhibition of the angiogenic response elicited by FGF2-T-MAE cells in a zebrafish/tumor xenograft model in which injection of mammalian tumor cells into the perivitelline space induces

Current Pharmaceutical Design, 2009, Vol. 15, No. 30

3583

the formation of tumor-driven neovessels sprouting from the sub-intestinal plexus of the embryo [105]. Accordingly, preliminary observations have shown that PTX3 overexpression decreases tumor growth and vascularization when PTX3 transfected murine melanoma B16/BL6 cells are embedded in Matrigel and grafted s.c. in syngeneic animals (Fig. 4). When considering the important cross-talk among innate immunity, inflammation, and tumor progression [106], these results raise the possibility that PTX3 may inhibit tumor growth and angiogenesis driven by FGF2. Relevant to this point, upregulation of PTX3 expression has been observed in human soft tissue liposarcoma [107] and in prostate cancer [108, 109]. It must be pointed out that PTX3 interacts with and inhibits the biological activity of FGF2 at doses comparable to those measured in the blood of patients affected by inflammatory diseases [86, 89]. Moreover, due to its capacity to accumulate in the ECM, the local concentration of PTX3 at the site of inflammation should be significantly higher than that measured in the blood stream, supporting the possibility that PTX3/FGF2 interaction may indeed occur and be biologically relevant in vivo. The putative effects of PTX3 on blood vessels are not limited to the angiogenic process during tumorigenesis. Indeed, PTX3 is present in atherosclerotic plaques [87] and is expressed by SMCs isolated from human arterial specimens [88]. Also, FGF2 is produced by SMCs and other cell types in the restenotic area, including ECs, macrophages, and T-cells [26]. Thus, FGF2 may exert autocrine and/or paracrine functions on FGFR-expressing SMCs of the injured vessel [22, 110]. In keeping with its FGF2 antagonist activity, PTX3 inhibits human coronary artery SMC prolife-

Fig. (4). PTX3 overexpression inhibits tumor growth and vascularization. Mock-transfected (upper panels) and PTX3-transfected (lower panels) murine melanoma B16/BL6 cells were implanted subcutaneously in syngeneic animals (3.0x105 cells in 400 μl of Matrigel). After 7 days, Matrigel plugs were harvested and snap-frozen. Sections were stained with haematoxylin and eosin (H&E) or double-immunostained with anti-CD31 and anti-PTX3 antibodies. Note the reduced cellularity and lack of CD31+ neovessels in PTX3-overexpressing plugs. Original magnification: x200.

3584 Current Pharmaceutical Design, 2009, Vol. 15, No. 30

ration and migration driven by endogenous and exogenous FGF2. Accordingly, PTX3 overexpression following recombinant adeno-associated virus gene transfer affects SMC proliferation and survival in vitro and intimal thickening after arterial injury in vivo [103]. 3.3. PTX3-Derived Antiangiogenic Peptides As described above, PTX3 binds FGF2 via its unique Nterminal extension [73]. To further define the FGF2-binding region in PTX3 N-terminus, the synthetic peptide PTX3(82110) and the overlapping peptides PTX3(82-96), PTX3(82101), and PTX3(97-110) were evaluated for their FGF2 antagonist activity together with three distinct peptides PTX3(31-60), PTX3(57-85), and PTX3(107-132) partially spanning the PTX3 N-terminus amino acid sequence (Fig. 5). The results demonstrated that the PTX3(82-110) peptide and the shorter PTX3(97-110) peptide show a similar FGF2 binding capacity and antagonist activity in vitro and in vivo, thus indicating that the amino acid linear sequence PTX3(97110) is responsible for PTX3/FGF2 interaction [73]. This region is predicted in an exposed loop of the PTX3 Nterminus that comprises the end of the B helix (Glu97), a turn on residues Ala104-Pro105-Gly106-Ala107, and the first two residues of the C helix (Ala109-Glu110) [73, 75]. On this basis, in an attempt to identify novel FGF2 antagonist(s), four acetylated (Ac) overlapping synthetic peptides based on the amino acid sequences PTX3(97-107), PTX3(100-113), PTX3(100-104), and PTX3(104-113) were compared with PTX3(97–110) peptide for their capacity to interact with FGF2 [111] (Fig. 5). Among them, the shortest pentapeptide Ac-ARPCA-NH2 [in single letter code, corresponding to amino acid sequence PTX3(100-104)] inhibits the

Leali et al.

interaction of FGF2 with PTX3 immobilized to a BIAcore sensorchip and suppresses FGF2-dependent proliferation in ECs. Also, Ac-ARPCA-NH2 inhibits angiogenesis triggered by FGF2 or by tumorigenic FGF2-T-MAE cells in chick and zebrafish embryos, respectively. Accordingly, the peptide hampers the binding of FGF2 to CHO cells overexpressing FGFR1 and to recombinant FGFR1 immobilized to a BIAcore sensorchip without affecting heparin interaction [111]. To assess the relevance of each amino acid residue of AcARPCA-NH2 pentapeptide for FGF2 interaction, a series of synthetic peptides carrying different amino acid substitutions were tested for their FGF2 antagonist activity. Scrambled Ac-PARAC-NH2 and Ac-APCRA-NH2 peptides showed a reduced inhibitory activity, pointing to the relevance of the relative position of each residue for the FGF2 antagonist capacity of the peptide. Also, the FGF2 antagonist activity was dramatically reduced for the non-acetylated H-ARPCANH2 peptide and for the Ac-ARPCG-NH2 and Ac-GRPCGNH2 peptides, but not for the Ac-GRPCA-NH2 peptide, indicating an essential role for the N-terminal blocking methyl group and for the methyl group of the side-chain of Ala5 residue in FGF2 interaction. The activity was lost also following amino acid substitution of residues Arg2, Pro3, or Cys4, thus underlying the role of the RPC amino acid sequence in Ac-ARPCA-NH2/FGF2 interaction. Interestingly, the FGF2 antagonist activity was lost also when the ARPCA region was mutated within synthetic peptides based on the amino acid sequences PTX3(97–110) and PTX3(82110), thus confirming the importance of the linear ARPCA sequence in PTX3/FGF2 interaction [111]. In order to map the peptide residues making direct contacts with FGF2, we applied Saturation Transfer Difference

Fig. (5). FGF2 antagonist activity of PTX3-derived peptides. Schematic representation of the synthetic peptides spanning the N-terminal domain of PTX3 and tested for their FGF2 antagonist activity [73, 98, 111]. Active peptide: grey bar; inactive peptide: open bar. All the active peptides contain the minimal linear FGF2-binding sequence PTX3(100-104) ARPCA (in black).

Antiangiogenic Activity of PTX3

(STD) NMR methods [112] to a series of Ac-ARPCA-NH 2 mutants in the presence of FGF2 [111]. In particular, the STD spectrum of Ac-ARPCA-NH2 peptide in the presence of FGF2 proves that the methyl protons of Ala1, Ala5, and of the N-terminal blocking acetyl group receive saturation transfer from the protein, indicating that these groups are the main responsible for the direct contact with FGF2. On the other hand, the RPC sequence may play a conformational role in ARPCA/FGF2 interaction, helping to orient the methyl groups of the peptide for optimal interaction with the growth factor. The extracellular portion of FGFRs comprises three Iglike domains (D1, D2, and D3, with an acidic box between D1 and D2). Their ligand binding and specificity reside in D2, D3, and D2-D3 linker region. X-ray crystallography has shown that the interactions between FGF2 and D3 are of both hydrophobic and polar character whereas the interactions with the D2-D3 linker are mediated mainly via hydrogen bonds. At variance, hydrophobic interactions dominate the interface between FGF2 and D2 [113]. Indeed, hydrophobic residues from discontinuous regions in FGF2, including Tyr24, Phe31, Tyr103, Leu140, and Met142, form a flat solvent-exposed hydrophobic surface which interacts hydrophobically with Leu165, Ala167, Pro169, and Val248 of the D2 domain in FGFR1. These residues are well conserved among the four mammalian FGFRs, indicating that this hydrophobic interface represents a highly conserved interaction site for FGF family members [114]. On this basis, Ac-ARPCA-NH2 peptide may exert its FGF2 antagonist activity by mimicking the hydrophobic ligand-binding region of D2, thus establishing hydrophobic interactions with the receptor-binding domain of FGF2 and competing with FGFRs for the binding to the growth factor. Accordingly, when a model of the proposed interaction was built by conformational analysis, the superposition of the global

Current Pharmaceutical Design, 2009, Vol. 15, No. 30

3585

minimum conformation of Ac-ARPCA-NH2 peptide to the -sheet region 164-170 of the hydrophobic domain D2 of FGFR1 indicated that the peptide could interact with FGF2 by mimicking this highly conserved FGF2-binding region of the receptor (Fig. 6). Hydrophobic interactions with FGF2 have been observed also for the C-terminal fragment of the FGF2 antagonist chemokine CXCL4 [115]. Moreover, an anti-angiogenic hydrophobic peptide has been identified by screening a phage display heptapeptide library following FGF2 biopanning [116]. At variance with Ac-ARPCA-NH2, this FGF2-binding peptide shares significant amino acid homology, charge distribution, and hydrophobic profile with the Ig-like domain D3 of FGFR1 and FGFR2. Thus, the complexity of FGF2/FGFR interaction is reflected by the possibility to generate various FGF2 antagonists endowed with the capacity to affect this interaction at different levels. 4. CONCLUDING REMARKS PTX3, a member of the pentraxin superfamily released locally by endothelial and inflammatory cells, binds the angiogenic growth factor FGF2 with high affinity and specificity, acting as a natural inhibitor of the autocrine and paracrine activity exerted by the growth factor on ECs. Thus, PTX3 may play an important role in modulating the crosstalk between inflammatory cells and endothelium in various physiopathological settings, including innate immunity, wound healing, restenosis, and atherosclerosis. Also, preliminary observations indicate that PTX3 overexpression may affect tumor growth via angiogenesis-dependent and independent mechanisms of action. Further experiments are required to clarify the impact of PTX3 on tumor progression and the possibility to design PTX3-derived anti-angiogenic and/or anti-neoplastic agents. To this respect, the identi-

Fig. (6). Hypothesis of interaction between Ac-ARPCA-NH2 and FGF2. The global minimum conformation of Ac-ARPCA-NH2 peptide (A) mimics the highly conserved -sheet portion 164-170 of the hydrophobic domain D2 of FGFR1 (B) interacting with FGF2 (PDB code: 1FQ9) [111]. Dotted spheres represent the hydrophobic groups of Ac-ARPCA-NH2 (N-terminal Ac, Ala1, and Ala5) and of FGFR1 (Leu165, Ala167, and Pro169) involved in the interaction with FGF2 (Tyr24, Phe31, Tyr103, Leu140, and Met142 in atom type).

3586 Current Pharmaceutical Design, 2009, Vol. 15, No. 30

fication of the acetylated Ac-ARPCA-NH2 pentapeptide as a short FGF2-binding peptide able to interfere with FGF2/ FGFR interaction and to exert a significant FGF2 antagonist activity may provide the basis for the design of novel PTX3derived peptidomimetic FGF2 antagonists.

Leali et al. [19]

[20]

ACKNOWLEDGEMENTS [21]

This work was supported by grants from Istituto Superiore di Sanità (Oncotechnological Program), Ministero dell’Istruzione, Università e Ricerca (Centro di Eccellenza per l’Innovazione Diagnostica e Terapeutica, Cofin projects), Associazione Italiana per la Ricerca sul Cancro, Fondazione Berlucchi, NOBEL Project Cariplo; and Fondazione Cariplo (Grant 2008-2264) to MP. PA is supported by a FIRC Fellowship. REFERENCES [2] [3] [4] [5] [6]

[7] [8]

[9]

[10]

[11]

[12] [13]

[14] [15]

[16]

[17]

[18]

Carmeliet P. Mechanisms of angiogenesis and arteriogenesis. Nat Med 2000; 6: 389-95. Hanahan D, Folkman J. Patterns and emerging mechanisms of the angiogenic switch during tumorigenesis. Cell 1996; 86: 353-64. Folkman J. Angiogenesis in cancer, vascular, rheumatoid and other disease. Nat Med 1995; 1: 27-31. Itoh N, Ornitz DM. Evolution of the Fgf and Fgfr gene families. Trends Genet 2004; 20: 563-9. Fantl WJ, Escobedo JA, Martin GA, Turck CW, Del Rosario M, McCormick F, et al. Distinct phosphotyrosines on a growth factor receptor bind to specific molecules that mediate different signaling pathways. Cell 1992; 69: 413-23. Javerzat S, Auguste P, Bikfalvi A. The role of fibroblast growth factors in vascular development. Trends Mol Med 2002; 8: 483-9. Dell'Era P, Belleri M, Stabile H, Massardi ML, Ribatti D, Presta M. Paracrine and autocrine effects of fibroblast growth factor-4 in endothelial cells. Oncogene 2001; 20: 2655-63. Redington AE, Roche WR, Madden J, Frew AJ, Djukanovic R, Holgate ST, et al. Basic fibroblast growth factor in asthma: measurement in bronchoalveolar lavage fluid basally and following allergen challenge. J Allergy Clin Immunol 2001; 107: 384-7. Kanazawa S, Tsunoda T, Onuma E, Majima T, Kagiyama M, Kikuchi K. VEGF, basic-FGF, and TGF-beta in Crohn's disease and ulcerative colitis: a novel mechanism of chronic intestinal inflammation. Am J Gastroenterol 2001; 96: 822-8. Yamashita A, Yonemitsu Y, Okano S, Nakagawa K, Nakashima Y, Irisa T, et al. Fibroblast growth factor-2 determines severity of joint disease in adjuvant-induced arthritis in rats. J Immunol 2002; 168: 450-7. Gibran NS, Isik FF, Heimbach DM, Gordon D. Basic fibroblast growth factor in the early human burn wound. J Surg Res 1994; 56: 226-34. Presta M, Dell'Era P, Mitola S, Moroni E, Ronca R, Rusnati M. Fibroblast growth factor/fibroblast growth factor receptor system in angiogenesis. Cytokine Growth Factor Rev 2005; 16: 159-78. Ornitz DM, Xu J, Colvin JS, McEwen DG, MacArthur CA, Coulier F, et al. Receptor specificity of the fibroblast growth factor family. J Biol Chem 1996; 271: 15292-7. Nicoli S, De Sena G, Presta M. Fibroblast Growth Factor 2-induced angiogenesis in zebrafish: the zebrafish yolk membrane (ZFYM) angiogenesis assay. J Cell Mol Med 2008; [Epub ahead of print]. Gualandris A, Rusnati M, Belleri M, Nelli EE, Bastaki M, Molinari-Tosatti MP, et al. Basic fibroblast growth factor overexpression in endothelial cells: an autocrine mechanism for angiogenesis and angioproliferative diseases. Cell Growth Differ 1996; 7: 147-60. Jackson JR, Seed MP, Kircher CH, Willoughby DA, Winkler JD. The codependence of angiogenesis and chronic inflammation. FASEB J 1997; 11: 457-65. Schmid MC, Varner JA. Myeloid cell trafficking and tumor angiogenesis. Cancer Lett 2007; 250: 1-8.

[22]

[23]

[24]

[25]

[26]

[27]

[28] [29]

[30] [31]

[32] [33]

[34] [35]

[36]

[37]

[38]

Kuwabara K, Ogawa S, Matsumoto M, Koga S, Clauss M, Pinsky DJ, et al. Hypoxia-mediated induction of acidic/basic fibroblast growth factor and platelet-derived growth factor in mononuclear phagocytes stimulates growth of hypoxic endothelial cells. Proc Natl Acad Sci USA 1995; 92: 4606-10. Baird A, Mormede P, Bohlen P. Immunoreactive fibroblast growth factor in cells of peritoneal exudate suggests its identity with macrophage-derived growth factor. Biochem Biophys Res Commun 1985; 126: 358-64. Blotnick S, Peoples GE, Freeman MR, Eberlein TJ, Klagsbrun M. T lymphocytes synthesize and export heparin-binding epidermal growth factor-like growth factor and basic fibroblast growth factor, mitogens for vascular cells and fibroblasts: differential production and release by CD4+ and CD8+ T cells. Proc Natl Acad Sci USA 1994; 91: 2890-94. Peoples GE, Blotnick S, Takahashi K, Freeman MR, Klagsbrun M, Eberlein TJ. T lymphocytes that infiltrate tumors and atherosclerotic plaques produce heparin-binding epidermal growth factor-like growth factor and basic fibroblast growth factor: a potential pathologic role. Proc Natl Acad Sci USA 1995; 92: 654751. Ribatti D, Crivellato E, Candussio L, Vacca A, Nico B, Benagiano V, et al. Angiogenic activity of rat mast cells in the chick embryo chorioallantoic membrane is down-regulated by treatment with recombinant human alpha-2a interferon and partly mediated by fibroblast growth factor-2. Haematologica 2002; 87: 465-71. Gloe T, Sohn HY, Meininger GA, Pohl U. Shear stress-induced release of basic fibroblast growth factor from endothelial cells is mediated by matrix interaction via integrin alpha(v)beta3. J Biol Chem 2002; 277: 23453-8. Hartnett ME, Garcia CM, D'Amore PA. Release of bFGF, an endothelial cell survival factor, by osmotic shock. Invest Ophthalmol Vis Sci 1999; 40: 2945-51. Cozzolino F, Torcia M, Lucibello M, Morbidelli L, Ziche M, Platt J, et al. Interferon-alpha and interleukin 2 synergistically enhance basic fibroblast growth factor synthesis and induce release, promoting endothelial cell growth. J Clin Invest 1993; 91: 2504-12. Lee HT, Lee JG, Na M, Kay EP. FGF-2 induced by interleukin-1 beta through the action of phosphatidylinositol 3-kinase mediates endothelial mesenchymal transformation in corneal endothelial cells. J Biol Chem 2004; 279: 32325-32. Walford G, Loscalzo J. Nitric oxide in vascular biology. J Thromb Haemost 2003; 1: 2112-8. Ziche M, Parenti A, Ledda F, Dell'Era P, Granger HJ, Maggi CA, et al. Nitric oxide promotes proliferation and plasminogen activator production by coronary venular endothelium through endogenous bFGF. Circ Res 1997; 80: 845-52. Finetti F, Solito R, Morbidelli L, Giachetti A, Ziche M, Donnini S. Prostaglandin E2 regulates angiogenesis via activation of fibroblast growth factor receptor-1. J Biol Chem 2008; 283: 2139-46. Gajdusek CM, Carbon S. Injury-induced release of basic fibroblast growth factor from bovine aortic endothelium. J Cell Physiol 1989; 139: 570-9. Sahni A, Altland OD, Francis CW. FGF-2 but not FGF-1 binds fibrin and supports prolonged endothelial cell growth. J Thromb Haemost 2003; 1: 1304-10. Sahni A, Francis CW. Stimulation of endothelial cell proliferation by FGF-2 in the presence of fibrinogen requires {alpha}v{beta}3. Blood 2004; 104: 3635-41. Wang L, Xiong M, Che D, Liu S, Hao C, Zheng X. The effect of hypoxia on expression of basic fibroblast growth factor in pulmonary vascular pericytes. J Tongji Med Univ 2000; 20: 265-7. Li J, Shworak NW, Simons M. Increased responsiveness of hypoxic endothelial cells to FGF2 is mediated by HIF-1alphadependent regulation of enzymes involved in synthesis of heparan sulfate FGF2-binding sites. J Cell Sci 2002; 115: 1951-9. Tiefenbacher CP, Chilian WM. Basic fibroblast growth factor and heparin influence coronary arteriolar tone by causing endotheliumdependent dilation. Cardiovasc Res 1997; 34: 411-7. Reuss B, Dono R, Unsicker K. Functions of fibroblast growth factor (FGF)-2 and FGF-5 in astroglial differentiation and bloodbrain barrier permeability: evidence from mouse mutants. J Neurosci 2003; 23: 6404-12. Zhang H, Issekutz AC. Growth factor regulation of neutrophilendothelial cell interactions. J Leukoc Biol 2001; 70: 225-32.

Antiangiogenic Activity of PTX3 [39]

[40]

[41]

[42] [43]

[44]

[45] [46] [47] [48]

[49]

[50] [51]

[52]

[53] [54]

[55]

[56]

[57]

[58] [59]

[60]

Barleon B, Sozzani S, Zhou D, Weich HA, Mantovani A, Marme D. Migration of human monocytes in response to vascular endothelial growth factor (VEGF) is mediated via the VEGF receptor flt-1. Blood 1996; 87: 3336-43. Wempe F, Lindner V, Augustin HG. Basic fibroblast growth factor (bFGF) regulates the expression of the CC chemokine monocyte chemoattractant protein-1 (MCP-1) in autocrine-activated endothelial cells. Arterioscler Thromb Vasc Biol 1997; 17: 2471-8. Jih YJ, Lien WH, Tsai WC, Yang GW, Li C, Wu LW. Distinct regulation of genes by bFGF and VEGF-A in endothelial cells. Angiogenesis 2001; 4: 313-21. Leali D, Dell'Era P, Stabile H, Sennino B, Chambers AF, Naldini A, et al. Osteopontin (Eta-1) and fibroblast growth factor-2 crosstalk in angiogenesis. J Immunol 2003; 171: 1085-93. Naldini A, Leali D, Pucci A, Morena E, Carraro F, Nico B, et al. Cutting edge: IL-1beta mediates the proangiogenic activity of osteopontin-activated human monocytes. J Immunol 2006; 177: 4267-70. Andres G, Leali D, Mitola S, Coltrini D, Camozzi M, Corsini M, et al. A Pro-inflammatory signature mediates FGF2-induced angiogenesis. J Cell Mol Med 2008; [Epub ahead of print]. Brower V. Antiangiogenesis research is booming, as questions and studies proliferate. J Natl Cancer Inst 2009; 101: 780-81. Casanovas O, Hicklin DJ, Bergers G, Hanahan D. Drug resistance by evasion of antiangiogenic targeting of VEGF signaling in latestage pancreatic islet tumors. Cancer Cell 2005; 8: 299-309. Brower V. How well do angiogenesis inhibitors work? biomarkers of response prove elusive. J Natl Cancer Inst 2009; 101: 846-47. Rusnati M, Presta M. Fibroblast growth factors/fibroblast growth factor receptors as targets for the development of anti-angiogenesis strategies. Curr Pharm Des 2007; 13: 2025-44. van Berkel V, Barrett J, Tiffany HL, Fremont DH, Murphy PM, McFadden G, et al. Identification of a gammaherpesvirus selective chemokine binding protein that inhibits chemokine action. J Virol 2000; 74: 6741-7. Parry CM, Simas JP, Smith VP, Stewart CA, Minson AC, Efstathiou S, et al. A broad spectrum secreted chemokine binding protein encoded by a herpesvirus. J Exp Med 2000; 191: 573-8. Fallon PG, Alcami A. Pathogen-derived immunomodulatory molecules: future immunotherapeutics? Trends Immunol 2006; 27: 470-6. Taraboletti G, Belotti D, Borsotti P, Vergani V, Rusnati M, Presta M, Giavazzi R. The 140-kilodalton antiangiogenic fragment of thrombospondin-1 binds to basic fibroblast growth factor. Cell Growth Differ 1997; 8: 471-9. Margosio B, Marchetti D, Vergani V, Giavazzi R, Rusnati M, Presta M, et al. Thrombospondin 1 as a scavenger for matrixassociated fibroblast growth factor 2. Blood 2003; 102: 4399-406. Bossard C, Van den Berghe L, Laurell H, Castano C, Cerutti M, Prats AC, et al. Antiangiogenic properties of fibstatin, an extracellular FGF-2-binding polypeptide. Cancer Res 2004; 64: 7507-12. Asplin IR, Wu SM, Mathew S, Bhattacharjee G, Pizzo SV. Differential regulation of the fibroblast growth factor (FGF) family by alpha(2)-macroglobulin: evidence for selective modulation of FGF-2-induced angiogenesis. Blood 2001; 97: 3450-7. Dennis PA, Saksela O, Harpel P, Rifkin DB. Alpha 2macroglobulin is a binding protein for basic fibroblast growth factor. J Biol Chem 1989; 264: 7210-6. De Marchis F, Ribatti D, Giampietri C, Lentini A, Faraone D, Scoccianti M, et al. Platelet-derived growth factor inhibits basic fibroblast growth factor angiogenic properties in vitro and in vivo through its alpha receptor. Blood 2002; 99: 2045-53. Perollet C, Han ZC, Savona C, Caen JP, Bikfalvi A. Platelet factor 4 modulates fibroblast growth factor 2 (FGF-2) activity and inhibits FGF-2 dimerization. Blood 1998; 91: 3289-99. Spinetti G, Camarda G, Bernardini G, Romano Di Peppe S, Capogrossi MC, Napolitano M. The chemokine CXCL13 (BCA-1) inhibits FGF-2 effects on endothelial cells. Biochem Biophys Res Commun 2001; 289: 19-24. Garlanda C, Bottazzi B, Bastone A, Mantovani A. Pentraxins at the crossroads between innate immunity, inflammation, matrix deposition, and female fertility. Annu Rev Immunol 2005; 23: 33766.

Current Pharmaceutical Design, 2009, Vol. 15, No. 30 [61]

[62] [63]

[64] [65] [66]

[67]

[68] [69]

[70] [71]

[72]

[73]

[74] [75]

[76]

[77]

[78]

[79]

[80]

[81]

[82]

3587

Goodman AR, Cardozo T, Abagyan R, Altmeyer A, Wisniewski HG, Vilcek J. Long pentraxins: an emerging group of proteins with diverse functions. Cytokine Growth Factor Rev 1996; 7: 191-202. Steel DM, Whitehead AS. The major acute phase reactants: Creactive protein, serum amyloid P component and serum amyloid A protein. Immunol Today 1994; 15: 81-8. Pepys MB, Baltz ML. Acute phase proteins with special reference to C-reactive protein and related proteins (pentaxins) and serum amyloid A protein. Adv Immunol 1983; 34: 141-212. Du Clos TW. The interaction of C-reactive protein and serum amyloid P component with nuclear antigens. Mol Biol Rep 1996; 23: 253-60. Gewurz H, Zhang XH, Lint TF. Structure and function of the pentraxins. Curr Opin Immunol 1995; 7: 54-64. Nauta AJ, Daha MR, van Kooten C, Roos A. Recognition and clearance of apoptotic cells: a role for complement and pentraxins. Trends Immunol 2003; 24: 148-54. Breviario F, d'Aniello EM, Golay J, Peri G, Bottazzi B, Bairoch A, et al. Interleukin-1-inducible genes in endothelial cells. Cloning of a new gene related to C-reactive protein and serum amyloid P component. J Biol Chem 1992; 267: 22190-7. Lee GW, Lee TH, Vilcek J. TSG-14, a tumor necrosis factor- and IL-1-inducible protein, is a novel member of the pentaxin family of acute phase proteins. J Immunol 1993; 150: 1804-12. Kretzschmar D, Hasan G, Sharma S, Heisenberg M, Benzer S. The swiss cheese mutant causes glial hyperwrapping and brain degeneration in Drosophila. J Neurosci 1997; 17: 7425-32. Lee GW, Goodman AR, Lee TH, Vilcek J. Relationship of TSG-14 protein to the pentraxin family of major acute phase proteins. J Immunol 1994; 153: 3700-7. Ortega-Hernandez OD, Bassi N, Shoenfeld Y, Anaya JM. The long pentraxin 3 and its role in autoimmunity. Semin Arthritis Rheum 2008; [Epub ahead of print]. Inforzato A, Peri G, Doni A, Garlanda C, Mantovani A, Bastone A, et al. Structure and function of the long pentraxin PTX3 glycosidic moiety: fine-tuning of the interaction with C1q and complement activation. Biochemistry 2006; 45: 11540-51. Camozzi M, Rusnati M, Bugatti A, Bottazzi B, Mantovani A, Bastone A, et al. Identification of an antiangiogenic FGF2-binding site in the N terminus of the soluble pattern recognition receptor PTX3. J Biol Chem 2006; 281: 22605-13. Berger B, Wilson DB, Wolf E, Tonchev T, Milla M, Kim PS. Predicting coiled coils by use of pairwise residue correlations. Proc Natl Acad Sci USA 1995; 92: 8259-63. Presta M, Camozzi M, Salvatori G, Rusnati M. Role of the soluble pattern recognition receptor PTX3 in vascular biology. J Cell Mol Med 2007; 11: 723-38. Inforzato A, Rivieccio V, Morreale AP, Bastone A, Salustri A, Scarchilli L, et al. Structural characterization of PTX3 disulfide bond network and its multimeric status in cumulus matrix organization. J Biol Chem 2008; 283: 10147-61. Introna M, Alles VV, Castellano M, Picardi G, De Gioia L, Bottazzai B, et al. Cloning of mouse ptx3, a new member of the pentraxin gene family expressed at extrahepatic sites. Blood 1996; 87: 1862-72. Bottazzi B, Vouret-Craviari V, Bastone A, De Gioia L, Matteucci C, Peri G, et al. Multimer formation and ligand recognition by the long pentraxin PTX3. Similarities and differences with the short pentraxins C-reactive protein and serum amyloid P component. J Biol Chem 1997; 272: 32817-23. Mantovani A, Garlanda C, Bottazzi B. Pentraxin 3, a nonredundant soluble pattern recognition receptor involved in innate immunity. Vaccine 2003; 21 (Suppl 2): S43-7. Vouret-Craviari V, Matteucci C, Peri G, Poli G, Introna M, Mantovani A. Expression of a long pentraxin, PTX3, by monocytes exposed to the mycobacterial cell wall component lipoarabinomannan. Infect Immun 1997; 65: 1345-50. Basile A, Sica A, d'Aniello E, Breviario F, Garrido G, Castellano M, et al. Characterization of the promoter for the human long pentraxin PTX3. Role of NF-kappaB in tumor necrosis factor-alpha and interleukin-1beta regulation. J Biol Chem 1997; 272: 8172-8. Altmeyer A, Klampfer L, Goodman AR, Vilcek J. Promoter structure and transcriptional activation of the murine TSG-14 gene encoding a tumor necrosis factor/interleukin-1-inducible pentraxin protein. J Biol Chem 1995; 270: 25584-90.

3588 Current Pharmaceutical Design, 2009, Vol. 15, No. 30 [83]

[84]

[85]

[86]

[87]

[88]

[89]

[90]

[91]

[92]

[93] [94] [95] [96]

[97]

[98]

[99] [100]

[101]

[102]

[103]

Polentarutti N, Bottazzi B, Di Santo E, Blasi E, Agnello D, Ghezzi P, et al. Inducible expression of the long pentraxin PTX3 in the central nervous system. J Neuroimmunol 2000; 106: 87-94. Doni A, Peri G, Chieppa M, Allavena P, Pasqualini F, Vago L, et al. Production of the soluble pattern recognition receptor PTX3 by myeloid, but not plasmacytoid, dendritic cells. Eur J Immunol 2003; 33: 2886-93. Abderrahim-Ferkoune A, Bezy O, Chiellini C, Maffei M, Grimaldi P, Bonino F, et al. Characterization of the long pentraxin PTX3 as a TNFalpha-induced secreted protein of adipose cells. J Lipid Res 2003; 44: 994-1000. Fazzini F, Peri G, Doni A, Dell'Antonio G, Dal Cin E, Bozzolo E, et al. PTX3 in small-vessel vasculitides: an independent indicator of disease activity produced at sites of inflammation. Arthritis Rheum 2001; 44: 2841-50. Rolph MS, Zimmer S, Bottazzi B, Garlanda C, Mantovani A, Hansson GK. Production of the long pentraxin PTX3 in advanced atherosclerotic plaques. Arterioscler Thromb Vasc Biol 2002; 22: e10-4. Klouche M, Peri G, Knabbe C, Eckstein HH, Schmid FX, Schmitz G, et al. Modified atherogenic lipoproteins induce expression of pentraxin-3 by human vascular smooth muscle cells. Atherosclerosis 2004; 175: 221-8. Peri G, Introna M, Corradi D, Iacuitti G, Signorini S, Avanzini F, et al. PTX3, A prototypical long pentraxin, is an early indicator of acute myocardial infarction in humans. Circulation 2000; 102: 63641. Luchetti MM, Piccinini G, Mantovani A, Peri G, Matteucci C, Pomponio G, et al. Expression and production of the long pentraxin PTX3 in rheumatoid arthritis (RA). Clin Exp Immunol 2000; 119: 196-202. Muller B, Peri G, Doni A, Torri V, Landmann R, Bottazzi B, et al. Circulating levels of the long pentraxin PTX3 correlate with severity of infection in critically ill patients. Crit Care Med 2001; 29: 1404-7. Salustri A, Garlanda C, Hirsch E, De Acetis M, Maccagno A, Bottazzi B, et al. PTX3 plays a key role in the organization of the cumulus oophorus extracellular matrix and in in vivo fertilization. Development 2004; 131: 1577-86. Milner CM, Day AJ. TSG-6: a multifunctional protein associated with inflammation. J Cell Sci 2003; 116: 1863-73. Milner CM, Higman VA, Day AJ. TSG-6: a pluripotent inflammatory mediator? Biochem Soc Trans 2006; 34: 446-50. Maina V, Cotena A, Doni A, Nebuloni M, Pasqualini F, Milner CM, et al. Coregulation in human leukocytes of the long pentraxin PTX3 and TSG-6. J Leukoc Biol 2009. Varani S, Elvin JA, Yan C, DeMayo J, DeMayo FJ, Horton HF, et al. Knockout of pentraxin 3, a downstream target of growth differentiation factor-9, causes female subfertility. Mol Endocrinol 2002; 16: 1154-67. Fulop C, Szanto S, Mukhopadhyay D, Bardos T, Kamath RV, Rugg MS, et al. Impaired cumulus mucification and female sterility in tumor necrosis factor-induced protein-6 deficient mice. Development 2003; 130: 2253-61. Rusnati M, Camozzi M, Moroni E, Bottazzi B, Peri G, Indraccolo S, et al. Selective recognition of fibroblast growth factor-2 by the long pentraxin PTX3 inhibits angiogenesis. Blood 2004; 104: 92-9. Johnson DE, Williams LT. Structural and functional diversity in the FGF receptor multigene family. Adv Cancer Res 1993; 60: 141. Schlessinger J, Lax I, Lemmon M. Regulation of growth factor activation by proteoglycans: what is the role of the low affinity receptors? Cell 1995; 83: 357-60. Richard C, Liuzzo JP, Moscatelli D. Fibroblast growth factor-2 can mediate cell attachment by linking receptors and heparan sulfate proteoglycans on neighboring cells. J Biol Chem 1995; 270: 2418896. Leali D, Belleri M, Urbinati C, Coltrini D, Oreste P, Zoppetti G, et al. Fibroblast growth factor-2 antagonist activity and angiostatic capacity of sulfated Escherichia coli K5 polysaccharide derivatives. J Biol Chem 2001; 276: 37900-8. Camozzi M, Zacchigna S, Rusnati M, Coltrini D, Ramirez-Correa G, Bottazzi B, et al. Pentraxin 3 inhibits fibroblast growth factor 2dependent activation of smooth muscle cells in vitro and neointima formation in vivo. Arterioscler Thromb Vasc Biol 2005; 25: 183742.

Leali et al. [104]

[105]

[106] [107]

[108]

[109]

[110] [111]

[112]

[113] [114]

[115]

[116] [117]

[118]

[119]

[120]

[121]

[122]

[123] [124]

Liekens S, Neyts J, De Clercq E, Verbeken E, Ribatti D, Presta M. Inhibition of fibroblast growth factor-2-induced vascular tumor formation by the acyclic nucleoside phosphonate cidofovir. Cancer Res 2001; 61: 5057-64. Nicoli S, Ribatti D, Cotelli F, Presta M. Mammalian tumor xenografts induce neovascularization in zebrafish embryos. Cancer Res 2007; 67: 2927-31. Lin WW, Karin M. A cytokine-mediated link between innate immunity, inflammation, and cancer. J Clin Invest 2007; 117: 1175-83. Willeke F, Assad A, Findeisen P, Schromm E, Grobholz R, von Gerstenbergk B, et al. Overexpression of a member of the pentraxin family (PTX3) in human soft tissue liposarcoma. Eur J Cancer 2006; 42: 2639-46. Sardana G, Jung K, Stephan C, Diamandis EP. Proteomic analysis of conditioned media from the PC3, LNCaP, and 22Rv1 prostate cancer cell lines: discovery and validation of candidate prostate cancer biomarkers. J Proteome Res 2008; 7: 3329-38. Ravenna L, Sale P, Di Vito M, Russo A, Salvatori L, Tafani M, et al. Up-regulation of the inflammatory-reparative phenotype in human prostate carcinoma. Prostate 2009; 69(1): 1245-55. Miyamoto T, Leconte I, Swain JL, Fox JC. Autocrine FGF signaling is required for vascular smooth muscle cell survival in vitro. J Cell Physiol 1998; 177: 58-67. Leali D, Bianchi R, Bugatti A, Nicoli S, Mitola S, Ragona L, et al. Fibroblast growth factor 2-antagonist activity of a long-pentraxin 3derived antiangiogenic pentapeptide. J Cell Mol Med 2009; [Epub ahead of print]. Mayer M, Meyer B. Group epitope mapping by saturation transfer difference NMR to identify segments of a ligand in direct contact with a protein receptor. J Am Chem Soc 2001; 123: 6108-17. Mohammadi M, Olsen SK, Ibrahimi OA. Structural basis for fibroblast growth factor receptor activation. Cytokine Growth Factor Rev 2005; 16: 107-37. Plotnikov AN, Hubbard SR, Schlessinger J, Mohammadi M. Crystal structures of two FGF-FGFR complexes reveal the determinants of ligand-receptor specificity. Cell 2000; 101: 413-24. Ragona L, Tomaselli S, Quemener C, Zetta L, Bikfalvi A. New insights into the molecular interaction of the C-terminal sequence of CXCL4 with fibroblast growth factor-2. Biochem Biophys Res Commun 2009; 382: 26-9. Wu X, Yan Q, Huang Y, Huang H, Su Z, Xiao J, et al. Isolation of a novel basic FGF-binding peptide with potent antiangiogenetic activity. J Cell Mol Med 2008; [Epub ahead of print].. Oyama S, Jr., Miranda MT, Toma IN, Viviani W, Gambarini AG. Mitogenic activity of peptides related to the sequence of human fibroblast growth factor-1. Biochem Mol Biol Int 1996; 39: 123744. Kiyota S, Franzoni L, Nicastro G, Benedetti A, Oyama S, Jr., Viviani W, et al. Introduction of a chemical constraint in a short peptide derived from human acidic fibroblast growth factor elicits mitogenic structural determinants. J Med Chem 2003; 46: 2325-33. Williams EJ, Williams G, Howell FV, Skaper SD, Walsh FS, Doherty P. Identification of an N-cadherin motif that can interact with the fibroblast growth factor receptor and is required for axonal growth. J Biol Chem 2001; 276: 43879-86. Facchiano A, Russo K, Facchiano AM, De Marchis F, Facchiano F, Ribatti D, et al. Identification of a novel domain of fibroblast growth factor 2 controlling its angiogenic properties. J Biol Chem 2003; 278: 8751-60. Presta M, Rusnati M, Urbinati C, Sommer A, Ragnotti G. Biologically active synthetic fragments of human basic fibroblast growth factor (bFGF): identification of two Asp-Gly-Argcontaining domains involved in the mitogenic activity of bFGF in endothelial cells. J Cell Physiol 1991; 149: 512-24. Terada T, Mizobata M, Kawakami S, Yabe Y, Yamashita F, Hashida M. Basic fibroblast growth factor-binding peptide as a novel targeting ligand of drug carrier to tumor cells. J Drug Target 2006; 14: 536-45. Ray J, Baird A, Gage FH. A 10-amino acid sequence of fibroblast growth factor 2 is sufficient for its mitogenic activity on neural progenitor cells. Proc Natl Acad Sci USA 1997; 94: 7047-52. Schubert D, Ling N, Baird A. Multiple influences of a heparinbinding growth factor on neuronal development. J Cell Biol 1987; 104: 635-43.

Antiangiogenic Activity of PTX3 [125]

[126] [127]

[128]

[129]

[130]

[131]

[132]

[133]

[134] [135]

[136] [137]

Current Pharmaceutical Design, 2009, Vol. 15, No. 30

Baird A, Schubert D, Ling N, Guillemin R. Receptor- and heparinbinding domains of basic fibroblast growth factor. Proc Natl Acad Sci USA 1988; 85: 2324-8. Walicke PA, Feige JJ, Baird A. Characterization of the neuronal receptor for basic fibroblast growth factor and comparison to receptors on mesenchymal cells. J Biol Chem 1989; 264: 4120-6. Lin X, Takahashi K, Campion SL, Liu Y, Gustavsen GG, Pena LA, et al. Synthetic peptide F2A4-K-NS mimics fibroblast growth factor-2 in vitro and is angiogenic in vivo. Int J Mol Med 2006; 17: 833-9. Cosic I, Drummond AE, Underwood JR, Hearn MT. In vitro inhibition of the actions of basic FGF by a novel 16 amino acid peptide. Mol Cell Biochem 1994; 130: 1-9. Kono K, Ueba T, Takahashi JA, Murai N, Hashimoto N, Myoumoto A, et al. In vitro growth suppression of human glioma cells by a 16-mer oligopeptide: a potential new treatment modality for malignant glioma. J Neurooncol 2003; 63: 163-71. Ito C, Saitoh Y, Fujita Y, Yamazaki Y, Imamura T, Oka S, et al. Decapeptide with fibroblast growth factor (FGF)-5 partial sequence inhibits hair growth suppressing activity of FGF-5. J Cell Physiol 2003; 197: 272-83. Li S, Christensen C, Kiselyov VV, Kohler LB, Bock E, Berezin V. Fibroblast growth factor-derived peptides: functional agonists of the fibroblast growth factor receptor. J Neurochem 2008; 104: 66782. Tzeng SF, Deibler GE, DeVries GH. Myelin basic protein and myelin basic protein peptides induce the proliferation of Schwann cells via ganglioside GM1 and the FGF receptor. Neurochem Res 1999; 24: 255-60. Kiselyov VV, Skladchikova G, Hinsby AM, Jensen PH, Kulahin N, Soroka V, et al. Structural basis for a direct interaction between FGFR1 and NCAM and evidence for a regulatory role of ATP. Structure 2003; 11: 691-701. Berezin V, Bock E. NCAM mimetic peptides: an update. Neurochem Res 2008; [Epub ahead of print]. Neiiendam JL, Kohler LB, Christensen C, Li S, Pedersen MV, Ditlevsen DK, et al. An NCAM-derived FGF-receptor agonist, the FGL-peptide, induces neurite outgrowth and neuronal survival in primary rat neurons. J Neurochem 2004; 91: 920-35. Hansen SM, Li S, Bock E, Berezin V. Synthetic NCAM-derived Ligands of the Fibroblast Growth Factor Receptor. Neurochem Res 2008. Anderson AA, Kendal CE, Garcia-Maya M, Kenny AV, MorrisTriggs SA, Wu T, et al. A peptide from the first fibronectin domain

Received: July 14, 2009

Accepted: July 16, 2009

[138]

[139]

[140]

[141]

[142]

[143]

[144]

[145] [146]

[147]

3589

of NCAM acts as an inverse agonist and stimulates FGF receptor activation, neurite outgrowth and survival. J Neurochem 2005; 95: 570-83. Jacobsen J, Kiselyov V, Bock E, Berezin V. A peptide motif from the second fibronectin module of the neural cell adhesion molecule, NCAM, NLIKQDDGGSPIRHY, is a binding site for the FGF receptor. Neurochem Res 2008; 33(12): 2532-9. Hansen SM, Kohler LB, Li S, Kiselyov V, Christensen C, Owczarek S, et al. NCAM-derived peptides function as agonists for the fibroblast growth factor receptor. J Neurochem 2008; 106: 2030-41. Hagedorn M, Zilberberg L, Lozano RM, Cuevas P, Canron X, Redondo-Horcajo M, et al. A short peptide domain of platelet factor 4 blocks angiogenic key events induced by FGF-2. Faseb J 2001; 15: 550-2. Yayon A, Aviezer D, Safran M, Gross JL, Heldman Y, Cabilly S, et al. Isolation of peptides that inhibit binding of basic fibroblast growth factor to its receptor from a random phage-epitope library. Proc Natl Acad Sci USA 1993; 90: 10643-7. Ballinger MD, Shyamala V, Forrest LD, Deuter-Reinhard M, Doyle LV, Wang JX, et al. Semirational design of a potent, artificial agonist of fibroblast growth factor receptors. Nat Biotechnol 1999; 17: 1199-204. Maruta F, Parker AL, Fisher KD, Hallissey MT, Ismail T, Rowlands DC, et al. Identification of FGF receptor-binding peptides for cancer gene therapy. Cancer Gene Ther 2002; 9: 54352. McConnell SJ, Thon VJ, Spinella DG. Isolation of fibroblast growth factor receptor binding sequences using evolved phage display libraries. Comb Chem High Throughput Screen 1999; 2: 155-63. Fan H, Duan Y, Zhou H, Li W, Li F, Guo L, et al. Selection of peptide ligands binding to fibroblast growth factor receptor 1. IUBMB Life 2002; 54: 67-72. Kanda S, Shono T, Tomasini-Johansson B, Klint P, Saito Y. Role of thrombospondin-1-derived peptide, 4N1K, in FGF-2-induced angiogenesis. Exp Cell Res 1999; 252: 262-72. Margosio B, Rusnati M, Bonezzi K, Cordes BL, Annis DS, Urbinati C, et al. Fibroblast growth factor-2 binding to the thrombospondin-1 type III repeats, a novel antiangiogenic domain. Int J Biochem Cell Biol 2008; 40: 700-9.