PDGF E. W. Raines* Department of Pathology, University of Washington, Health Science Building J507 Box 357470, Seattle, WA 98195-7470, USA * corresponding author tel: 206-685-7441, fax: 206-685-3018, e-mail:
[email protected] DOI: 10.1006/rwcy.2000.08003.
SUMMARY Different dimeric forms of platelet-derived growth factor (PDGF) are key regulators of connective tissue cells in embryogenesis and in the pathogenesis of a number of diseases. During development, PDGF-B/ PDGF receptor signaling controls mesangial cell development in kidney glomeruli (LeveÂen et al., 1994; Lindahl et al., 1998) and the recruitment of vascular smooth muscle cells (SMCs) and pericytes in developing blood vessels (Lindahl et al., 1997a; Crosby et al., 1998; HellstroÈm et al., 1999), while PDGF-A/ PDGF receptor signaling is critical for the recruitment of alveolar SMCs and alveogenesis (BostroÈm et al., 1996; Lindahl et al., 1997b) and for O-2A progenitor cell expansion in developing rat optic nerve, which gives rise to oligodendrocytes and type 2 astrocytes (Noble et al., 1988; Raff et al., 1988). The newly defined PDGF-C also targets mesenchymal cells, and may be particularly important in the development of the kidney mesenchyme and physiological and pathophysiological remodeling of the cardiac interstitium (Li et al., 2000). The potent effects of PDGF on mesenchymal cells in development, particularly SMCs, is recapitulated in wound healing and in multiple inflammatory diseases, where PDGF appears to be important for the recruitment of connective tissue cells and matrix deposition in the fibrotic response.
BACKGROUND
Discovery PDGF was originally discovered because it was the principal source of growth factor activity in whole blood serum for mesenchymal cells that was missing
in cell-free, plasma-derived serum (Kohler and Lipton, 1974; Ross et al., 1974). Although the possibility that platelets might serve as a source of growth factor activity was postulated by Balk (1971), it was not until it was shown that platelets could restore all of the growth factor activity missing from plasma that the significance of the platelet as a source of this activity became clear. These observations led to the purification and characterization of multiple molecular forms of PDGF from human platelets (Antoniades, 1981; Heldin et al., 1981a) and plateletrich plasma (Deuel et al., 1981; Raines and Ross, 1982).
Alternative names Because of the near identity of p28sis, the transforming product of simian sarcoma virus, with the B chain of PDGF (Robbins et al., 1983), PDGF is sometimes referred to as c-sis.
Structure PDGF is a family of homo- and heterodimers of two distinct cationic polypeptide chains (PDGF A chain and B chain), whose interchain and intrachain disulfide bonds (Giese et al., 1987; Hannink and Donoghue, 1988; OÈstman et al., 1992; Haniu et al., 1993) are responsible for its tight knot-like structure (Oefner et al., 1992). Recently, a third PDGF family member (PDGF-C) has been identified that only forms homodimers (Li et al., 2000).
Main activities and pathophysiological roles The ability of PDGF to stimulate mesenchymal migration, proliferation, survival, and multiple
756 E. W. Raines specific cellular functions, including extracellular matrix deposition, led to the hypothesis that it may be important for multiple fibroproliferative responses in disease states (Ross et al., 1974; Westermark and Wasteson, 1976; Ross et al., 1986). The deletion of either the PDGF A or B chain or either of the two PDGF receptors is embryonic lethal (LeveÂen et al., 1994; Soriano, 1994, 1997; BostroÈm et al., 1996). An analysis of the null embryos has demonstrated an important role for PDGF in mesenchymal cell recruitment in the kidney, blood vessels, lungs, and central nervous system. A unique role for PDGF-C is indicated from analysis of kidneys lacking the PDGF receptor that shows selective loss of mesenchymal cells adjacent to sites of PDGF-C mRNA expression that are not found in kidneys from animals lacking PDGF-A or both PDGF-A and PDGF-B (Li et al., 2000). Similarly, in wound repair and in multiple diseases, the PDGF recruitment of connective tissue cells contributes to the fibroproliferative response and to disease pathogenesis (Betsholtz and Raines, 1997).
GENE AND GENE REGULATION
Accession numbers PDGF A chain: NM 002607 (short); X03795 (long) PDGF B chain: Z81010 PDGF C chain: W21436 (EST)
Chromosome location The genes for the PDGF A and B chains are located on chromosomes 7 (Dalla Favera et al., 1982; Swan et al., 1982; Betsholtz et al., 1986) and 22 (Bonthron et al., 1988; Stenman et al., 1988) respectively.
Regulatory sites and corresponding transcription factors PDGF A Chain An unusual feature of the PDGF A chain mRNA is the presence of a long 50 untranslated region (UTR) containing three AUG triplets upstream of the initiator codon (Bonthron et al., 1988; Rorsman et al., 1988). The first two AUG are in close proximity in the same reading frame and are closely followed by a stop codon. It would be predicted that the complicated leader sequence might impair translation.
A highly GC-rich region, containing three contiguous specificity protein 1 (SP-1)-binding sites between bp ÿ150 and ÿ33, and overlapping early growth response factor 1 (Egr-1)-binding sites, are responsible for over 80% of the promoter activity (Lin et al., 1992; Wang and Deuel, 1992; Kaetzel et al., 1993). In smooth muscle cells, SP-1 and SP-3 bind to the promoter and independently or additively activate it (Silverman et al., 1997). Egr-1 is induced by a number of stimuli, including shear stress (Khachigian et al., 1997), and can compete with SP-1 to activate the promoter further (Silverman et al., 1997). Strainresponsive regions between ÿ92 bp and ÿ41 bp of the promoter have also been identified (Wilson et al., 1998), and strain only induced binding with an Egr-1 probe but not with an SP-1 probe. In contrast, methylation of the promoter appears to repress its activity (Lin et al., 1993). Regions involved in the repression of the PDGF A chain include those between ÿ1029 and ÿ883 and between ÿ1800 and ÿ1029 in the renal epithelial cell line BSC-1 (Kaetzel et al., 1994; Liu et al., 1996), and the first intron in HeLa cells (Wang et al., 1994; Nobuyoshi et al., 1997). The Wilms' tumor suppressor gene product binds to several sites in the PDGF A chain promoter and functions as a repressor (Wang et al., 1992; Lee et al., 1997). Recently, a cloned DNA binding factor, GC factor 2 (GCF2), has been shown to repress the human PDGF A chain promoter by competing with SP-1 and Egr-1 for interaction with the promoter (Khachigian et al., 1999). GCF2 is induced on the mechanical injury of cells in culture and after balloon injury of the rat carotid artery. Data are consistent with GCF2 acting as an endogenous transcriptional repressor. PDGF B Chain A minimal promoter for the PDGF B chain extending only 42 bp upstream of the TATA signal is as efficient as 4 kb upstream of the TATA in driving expression in unstimulated K562 cells (Pech et al., 1989). Laminar shear stress induces PDGF B chain expression, and a 6 bp core element has been defined as a shear stress response element (Resnick et al., 1993). Using electrophoretic mobility shift assays and in vitro DNAse I footprinting, it has been demonstrated that NFB p50±p65 heterodimers bind to the PDGF B chain shear stress response element in endothelial cells (Khachigian et al., 1995). Egr-1 has also been shown to interact with a novel element in the proximal PDGF B chain promoter and to compete for SP-1 binding after injury of endothelial cells (Khachigian et al., 1995). Elements associated with thrombin-induced PDGF B chain
PDGF 757 transcription in endothelial cells (Scarpati and DiCorleto, 1996) and the TPA induction of K562 cells (Pech et al., 1989) have been defined. In osteosarcoma cells, SP-1 and SP-3 have been identified as primary transcriptional mediators (Liang et al., 1996). SP-1 and SP-3 have also been shown to be elevated in newborn rat smooth muscle cells that constitutively express the PDGF B chain (Rafty and Khachigian, 1998). The long 50 UTR of the PDGF B chain strongly inhibits its synthesis (Horvath et al., 1995). Inaccessibility of the promoter, as a result of cell type-specific DNase-I hypersensitivity sites, appears to contribute to inhibition of the expression of the endogenous gene (Franklin et al., 1991; Dirks et al., 1993). Transactivating factors have also been implicated in the negative regulation of the PDGF B chain in somatic cell hybrids between PDGF B chain-expressing melanoma cells and hamster fibroblasts (LeveÂen et al., 1993). Furthermore, the long 50 UTR of the PDGF B chain has translational modulating activity due to its differentiation-activated internal ribosomal entry site, and the elements required for differentiation-sensing ability have been mapped to a 630 nucleotide fragment within the central portion of the 50 UTR (Bernstein et al., 1997; Sella et al., 1999). Transcripts lacking the translational inhibitory effect of exon 1 appear to be initiated within the first alternative exon (Fen and Daniel, 1991; Dirks et al., 1995). The expression of a 2.6 kb PDGF B chain mRNA lacking the inhibitory 50 untranslated sequence was found to correlate with increased PDGF B chain immunoreactivity during rat brain development (Sasahara et al., 1998).
Cells and tissues that express the gene PDGF may be expressed by a multiplicity of normal cells and tissues (Table 1). Although PDGF A, B, and C chains are expressed constitutively in the brain (Sasahara et al., 1991; Yeh et al., 1991; Li et al., 2000), their expression in normal cells is highly regulated (Raines et al., 1990; Raines and Ross, 1993). For a number of cell types, placing isolated cells in culture is sufficient to induce PDGF mRNA expression (Raines and Ross, 1993). Conversely, culture conditions that more closely mimic those in vivo decrease PDGF expression (Raines and Ross, 1993). PDGF expression in vivo is increased following injury and in a number of disease states (Table 2), including tumors of multiple cell origin (Table 3).
PROTEIN
Accession numbers GenPept: A chain: NP 002598 (short); CAA27421 (long) B chain: CAV02635 Crystal structure: PDGF-BB ± MMDB, 1872; PDB, 1PDG
Sequence The amino acid sequences for the PDGF A and B chains are shown in Figure 1a. The sequences have been aligned to allow a comparison of the amino acid sequences of the two chains, and demonstrate the significant homology between the two chains (51% over the 109 amino acids of the PDGF B chain). The newly cloned PDGF-C shares 27±35% homology with PDGF-A and PDGF-B over exons 4 through 7 (Li et al., 2000). Phylogenetic analysis of PDGF-C and domains of PDGF-A and -B and the related family of vascular endothelial growth factor (VEGF) shows that PDGF-C is closer to VEGFs than to PDGFs (Li et al., 2000).
Description of protein PDGF is a family of homo- and heterodimers of two distinct but homologous genes, PDGF A chain and B chain and homodimers of PDGF-C (see Figure 1). The processed forms of PDGF A, B, and C chain vary from 125, 108, and up to 184 amino acids in length, respectively. An analysis of PDGF purified from lysed human platelets, using chain-specific monoclonal antibodies, suggests that approximately 70% is PDGF-AB, the remainder being PDGF-AA and PDGF-BB (Hammacher et al., 1988a; Hart et al., 1990). Other cells that express both chains of PDGF also contain all three PDGF isoforms (Hammacher et al., 1988b; OÈstman et al., 1988), suggesting that dimer assembly may be a random process. The maintenance of the tertiary structure by intrachain disulfide bonds is required for biological activity (Heldin et al., 1979; Raines and Ross, 1982; Antoniades and Williams, 1983; Giese et al., 1987; Kenney et al., 1994). Both A and B chains are synthesized as precursor proteins (Figure 1b) and undergo processing at the Ntermini (A and B chain) and C-terminus (B chain). Two forms of the PDGF A chain are synthesized from alternatively spliced forms, with and without the
758 E. W. Raines Table 1 Normal cells and tissues that express PDGF Cell
Protein
mRNA
Reference
PDGF-A
PDGF-B
Connective tissue cells Fibroblasts Dermal, human
+
+
ÿ
Raines et al., 1989
Foreskin, human
+
+
ÿ
Paulsson et al., 1987
+
+
+
Gnessi et al., 1992, 1995
+
+
+
Marra et al., 1994
+
+
+
Shultz et al., 1988
+
+
ÿ
Sejersen et al., 1986
+
+
+
Gnessi et al., 1995
Aortic, carotid, femoral, human
+
+
ÿ
Libby et al., 1988
Aortic, adult rat
+
+
ÿ
Sejersen et al., 1986
Aortic, newborn rat
+
+
+
Seifert et al., 1984
Aortic, bovine
+
+
+
DiCorleto and Bowen-Pope, 1983
Iliac, human
+
+
+
Sitaras et al., 1987
Renal microvascular, human
+
+
+
Daniel et al., 1986
Umbilical vein, human
+
+
+
Collins et al., 1987a, 1987b
Venous, human
+
+
+
Limanni et al., 1988
+
+
+
Kim et al., 1999
Leydig cells Leydig cells, rat Liver fat-storing cells, human Mesangial cells Mesangial cells, human Muscle cells Myoblasts, rat skeletal Peritubular myoid cells Developing and postnatal, rat Vascular smooth muscle
Endothelial/epithelial cells Endothelial cells
Epithelial cells Corneal, human Iris pigment, human
+
+
Kociok et al., 1998
Keratinocytes, human
+
+
+
Ansel et al., 1993
Kidney, African green monkey (BSC-1)
+
ÿ
+
Kartha et al., 1988
Pancreatic islet cells, human
+
+
Retinal pigment, rat
+
+
+
Campochiaro et al., 1989; Kociok et al., 1998
Sertoli cells, developing rat
+
+
+
Gnessi et al., 1995
Ebert et al., 1998
Mesothelial cells Pleural, human
+
Gerwin et al., 1987
Hematopoietic cells Erythrocytes Erythroid cells, mouse
+ +
Keutzer and Sytkowski, 1995 +
+
Bidwell et al., 1995
PDGF 759 Table 1 (Continued ) Cell
Protein
mRNA PDGF-A
Reference PDGF-B
Monocyte/macrophages Alveolar, human
+
Blood monocytes, human
ÿ
Activated monocytes, human
+
+
+
Sariban and Kufe, 1988; Ross et al., 1990
+
+
+
Hart et al., 1990; Wickenhauser et al., 1995
+
+
+
Noble et al., 1988; Raff et al., 1988; Richardson et al., 1988
+
+
+
Sasahara et al., 1991
Platelets/megakaryocytes
+
Shimokado et al., 1985 Ross et al., 1990
Nervous tissue cells Type I astrocytes, rat Neurons Primate Mouse Retinal ganglion, rat Schwann cells, rat
+
Yeh et al., 1991
+
+
Fruttiger et al., 1996
+
+
+
Hardy et al., 1992
+
+
Barrett and Benditt, 1988
Tissues Aorta Human Bone, bovine
+
Hauschka et al., 1988
Brain Caudate-putamen, primate
+
Sasahara et al., 1991
Cerebellum, primate
+
+
+
Sasahara et al., 1991; Yeh et al., 1991
Cortex, primate
+
+
+
Sasahara et al., 1991; Yeh et al., 1991
Hippocampus, primate
+
+
+
Sasahara et al., 1991; Yeh et al., 1991
Optic nerve, rat
+
+
Sciatic nerve, rat
+
+
+
Hardy et al., 1992
Carotid endarterectomy, human
+
+
Barrett and Benditt, 1988
Cornea, human
+
+
Kim et al., 1999
Mudhar et al., 1993
Embryo Unfertilized oocyte, mouse Blastocyst, mouse
+
Xenopus Epidermis
+
Rappolee et al., 1988
+
Rappolee et al., 1988
+
Mercola et al., 1988
+
+
+
+
+
Ansel et al., 1993
Kidney Human Developing, mouse
+
Young et al., 1990; Alpers et al., 1995 +
Seifert et al., 1998
Liver, human
+
Young et al., 1990
Lung, human
+
Young et al., 1990
+
Young et al., 1990
Muscle Muscle, human
760 E. W. Raines Table 1 (Continued ) Cell
Protein
mRNA PDGF-A
Developing muscle, rat Pancreas, human
+
Reference PDGF-B
+
Jin et al., 1990
+
Ebert et al., 1998
Placenta First trimester, human
+
Second trimester, human
+
Third trimester, human
+
+
Retina, rat
+
+
Hyaloid rat
+
+
Goustin et al., 1985
+
Taylor and Williams, 1988 Gurski et al., 1999
Eye Mudhar et al., 1993 +
Mudhar et al., 1993
Thyroid, human
+
Young et al., 1990
Teeth, developing mouse
+
Chai et al., 1998
+
Young et al., 1990
Testis Testis, human Developing and postnatal, rat
+
+
+
Gnessi et al., 1995
Uterus, human
+
Boehm et al., 1990
Uterus, human gravid
+
Mendoza et al., 1990
Uterus
Table 2 PDGF ligand and receptor expression are increased in human disease Disease
PDGF chains A
Atherosclerosis
PDGF receptors B
+
Libby et al., 1988 +
+ Restenosis Cirrhosis of the liver
Ross et al., 1990
+
Wilcox et al., 1988
+ +
Reference
+
+
+
Rubin et al., 1988
+
Tanizawa et al., 1996
+
Pinzani et al., 1996
Diabetes
Guillausseau et al., 1989
Glomerulonephritis
+ +
Nakashima et al., 1992
+
Matsuda et al., 1997
+
Taniguchi et al., 1996 +
Inflammatory bowel disease +
+
+
+
FellstroÈm et al., 1989
+
Gesualdo et al., 1994
+
Alexander et al., 1995 Beck and Podolsky, 1999
PDGF 761 Table 2 (Continued ) Disease
PDGF chains
PDGF receptors
+
+
Reference
A
B
Myelofibrosis
+
+
Proliferative retinal diseases
+
+
Idiopathic
+
+
Nagaoka et al., 1990
Histiocytosis X
+
+
Uebelhoer et al., 1995
Katoh et al., 1990 Robbins et al., 1994
Pulmonary fibrosis
Rheumatoid arthritis
+
Reuterdahl et al., 1991
Scleroderma
+
Klareskog et al., 1990
+ Cardiac transplant rejection Renal transplant rejection
+ +
Gay et al., 1989 +
Zhao et al., 1995
+
Shaddy et al., 1996
+
Alpers et al., 1996
+
+
Floege et al., 1998 +
FellstroÈm et al., 1989
Table 3 Expression of PDGF by tumors and tumor cells Cell
Protein
mRNA
Reference
PDGF-A
PDGF-B
Connective tissue tumors Fibrosarcoma, human
+
+
+
Taniuchi et al., 1997
Kaposi's sarcoma, human
+
+
+
Sturzl et al., 1992
Osteosarcoma, human (U-2OS)
+
+
+
NisteÂr et al., 1988
+
+
Collins et al., 1987a
+
+
Raines et al., 1990
(U-1810) Epithelial cell tumors (carcinomas) Bladder, human (T24) Breast, human
+
+
Coltrera et al., 1995
+
+
+
Bronzert et al., 1987
Colon, human (COLO-201, COLO-205)
+
+
+
Sariban et al., 1988
Esophageal, human
+
+
+
Yoshida et al., 1993
Gastric, human
+
+
Chung and Antoniades, 1992
(KATO III)
+
+
Sariban et al., 1988
(MCF-7, MDA-MB-231)
Liver, human (Hep G2) Lung, human (CALU-1) (U-1810) Mammary, human (MDA-MB-468) (BT-20, MCF-7, ZR-75-1)
+
+
+
Bowen-Pope, 1984b
+
+
+
Raines et al., 1990
+
+
+
Sariban et al., 1988
+
+
+
Betsholtz et al., 1987
+
ÿ
+
Peres et al., 1987
+
+
+
Sariban et al., 1988
762 E. W. Raines Table 3 (Continued ) Cell
Protein
mRNA
Reference
PDGF-A
PDGF-B
+
+
+
Peres et al., 1987; Sariban et al., 1988
+
+
+
Barnhill et al., 1996
(WM-115)
+
+
+
Westermark et al., 1986
(WM-239A, WM-266-4)
+
+
ÿ
Westermark et al., 1986
+
+
+
Gerwin et al., 1987
+
+
Gerwin et al., 1987
+
ÿ
Gerwin et al., 1987
(MDA-MB-157, T47D, HBL-100) Melanoma, human
Mesothelial, human (HUT-28) (DND, JMN, MT-1, MT-3, VMAT-1) (HUT-290) Ovarian, human (ARM, DUN, MAC, SAM)
+
+
+
Sariban et al., 1988
Pancreatic, human (PANC-I and HPAF)
+
+
+
Ebert et al., 1995
Pituitary, human
+
+
+
Leon et al., 1994
Prostate, human adenocarcinomas
+
+
ÿ
Fudge et al., 1994
Erythroleukemic cells, human (HEL, K562)
+
+
+
Papayannopoulou et al., 1987
Pre-B leukemic cells, human (SMS-SB)
+
+
Hematopoietic cell tumors Tsai et al., 1994
Promyelocytic, human (HL-60)
+
Promonocytic, human (U937)
+
Alitalo et al., 1987
Promonocytic, human (THP-1)
+
Sariban and Kufe, 1988
T cells transformed with HTLV-I and II
+
Pantazis et al., 1986; Alitalo et al., 1987
+
+
+
Pantazis et al., 1987; Goustin et al., 1990
+
+
ÿ
Fraizer et al., 1987
(human stem cell line Tera-2 clone 13)
+
+
ÿ
Weima et al., 1988
(mouse endoderm-like F9)
+
+
ÿ
Wang and Stiles, 1993
+
Fahrer et al., 1989
+
+
Maxwell et al., 1990
+
+
Black et al., 1996
Mixed cell tumors Nephroblastoma, human Wilms' tumor Teratocarcinoma
Muscle cell tumors Rhabdomyosarcoma, human
+
Nervous tissue tumors Astrocytomas, human
+
Ependymomas, human Glioblastoma, human
+
ÿ
+
Gillaspy et al., 1992
(A172)
+
ÿ
+
Vassbotn et al., 1994
(patient with LiFraumeni syndrome)
+
+
ÿ
Guha et al., 1995
+
NisteÂr et al., 1988
Gliomas, human Medulloblastomas, human
+
ÿ
Black et al., 1996
Meningiomas, human
+
+
Black et al., 1994
+
Van Zoelen et al., 1985
+
Di Rocco et al., 1998
Neuroblastomas (mouse 2A) Oligodendrogliomas, human
+ +
PDGF 763 Figure 1 PDGF A- and B-chain: Amino acid sequences, protein processing, and different dimeric isoforms (a) The amino acid sequences of the PDGF A- and B-chain deduced from the nucleotide sequence of cloned cDNA are shown and aligned for comparison of homology between the two chains. The alignment is adapted from Bonthron et al. (1988) and dots represent gaps introduced to improve alignment. The N- and C-terminals cleavage sites are shown by arrows (A-chain, green; B-chain, blue). Boxes indicate exons and each exon is numbered above. Astrisks indicate matching residues. The two alternatively spliced sequences for the PDGF A-chain are shown for exons 6 and 7 (AL, the long from of the A-chain, yellow; AS, the short form of the A-chain, red). Figure adapted from Raines et al. (1990). (b) The processing of the PDGF A- and Bchains is illustrated with the signal sequence indicated by the zig-zag line, the pro-sequence by a line, and the mature protein is boxed (A-chain, green; and alternatively spliced sequences as indicated in (a)). Positions with consensus N-linked glycosylation sites have been indicated with a branched structure. (c) The six different dimeric combination of PDGF A- and B-chains are shown. (Full colour figure can be viewed online.)
exon 6-encoded sequence (Betsholtz et al., 1986; Collins et al., 1987a). In both the A and the B chain, exon 1 encodes the signal sequence, while exons 2 and 3 encode precursor sequences removed during processing, exons 4 and 5 encode the mature protein, exon 6 is alternatively spliced in the A chain and may be removed during maturation of the B chain, and exon 7 is primarily a noncoding sequence (Johnsson
et al., 1984; Bonthron et al., 1988; OÈstman et al., 1988, 1992; Rorsman et al., 1988). PDGF-C has a twodomain structure not previously observed in this family of growth factors: (1) an N-terminal CUB domain first found in complement subcomponents C1r/C1s, urchin EGF-like protein and bone morphogenetic protein 1, which may be involved in protein±protein and protein±carbohydrate interactions; (2) a C-terminal
764 E. W. Raines PDGF/VEGF-homology domain. The two domains are connected by a hinge region (Li et al., 2000). PDGF-C is synthesized and secreted as a latent growth factor, requiring proteolytic removal of the Nterminal CUB domain for receptor binding and activation. Alternative splicing of the PDGF A chain and processing of the C-terminus of the PDGF B chain regulate the inclusion or exclusion of the highly basic sequence encoded by exon 6 whose inclusion may promote intracellular retention (LaRochelle et al., 1991) and matrix association (Pollock and Richardson, 1992; Raines and Ross, 1992; Kelly et al., 1993; Andersson et al., 1994). If the A and B chains are made by a single cell and both alternatively spliced forms of the A chain are transcribed, six dimeric combinations are possible (Figure 1c). The recent expression of the full-length PDGF B chain and a truncated form lacking exon 6 in genetically engineered human keratinocytes that were grafted as epithelial sheets onto athymic mice demonstrated that the different binding properties control the spatial organization of cellular events in regenerating mesenchymal tissues (Eming et al., 1999).
Discussion of crystal structure The crystal structure of human recombinant PDGFBB (Oefner et al., 1992) demonstrates that the polypeptide chain is folded into two highly twisted, antiparallel pairs of strands. Each of the chains creates three loops, two (loops 1 and 3) extending in one direction and one (loop 2) in the opposite direction, which upon dimerization forms a complex of all three loops contributed to by different chains. Mutation of the loop 2 region of the PDGF A and B chain significantly decreases binding to both the PDGF and receptors (Andersson et al., 1995), while specific amino acid residues in loop 1 of the PDGF B chain appear to be important primarily for binding to the PDGF receptor; the Lys161 in loop 3 is involved in binding to the PDGF receptor (Kreysing et al., 1996; Schilling et al., 1998), and the three cationic amino acid residues in loop 3 also contribute to heparin binding activity (Schilling et al., 1998). The tight cystine-knot topology of PDGF chains is shared with not only the highly homologous VEGF family (Muller et al., 1997), but also two other growth factors, TGF and nerve growth factor (Murray-Rust et al., 1993). It has been suggested that the structural motif of disulfide bonds and hydrogen-bonded strands found in these growth factors with divergent amino acid sequences serves as a framework for the
elaboration of loops that contain the specificity for receptor interaction. In spite of the diversity in the sequence of PDGF-C with that of PDGF-A and -B, which includes the different spacing of the cysteine residues in the central, most highly conserved region of this domain with an insertion of three amino acids that occurs close to the loop 2 region thought to be involved in receptor binding, PDGF-C dimers bind to the PDGF receptor with an almost identical affinity to that of PDGF-AA or PDGF-BB (Li et al., 2000).
Important homologies PDGF A and B chains share 60% amino acid homology, with a perfect conservation of the eight cysteine residues. Members of the VEGF family share a similar spacing between cysteine residues (Keck et al., 1989; Tischer et al., 1989; Conn et al., 1990), although their target cells and activities are distinct. The eight conserved cysteine residues are also found in PDGF-C, but their spacing is different with the insertion of three amino acids between cysteines 3 and 4 (Li et al., 2000). An additional four cysteine residues are present in PDGF-C and are located between cysteines 3 and 4, 5 and 6, 6, and 7, and beyond the eighth conserved cysteine (Li et al., 2000). The overall sequence homology of PDGF-C with PDGF-A and -B is only 27 to 35%. The unique N-terminal CUB domain only found in PDGF-C shares 27±37% identity with the prototypic CUB domains in C1r/C1s and bone morphogenic protein 1 (Li et al., 2000).
Posttranslational modifications An analysis of processing of the v-sis gene product (Hannink and Donoghue, 1986) and the PDGF B chain (Kaetzel et al., 1996) demonstrates that mutants lacking the N-linked glycosylation site are properly folded, are secreted and are biologically active. However, in the case of the PDGF B chain, the absence of the N-glycosylation site decreases the accumulation of the intracellular species (Kaetzel et al., 1996).
CELLULAR SOURCES AND TISSUE EXPRESSION
Cellular sources that produce Although the PDGF B chain can be expressed in a number of normal cells in culture (see Table 1), its constitutive expression in normal cells and tissues
PDGF 765 appears to be primarily limited to extensive areas of the central nervous system (Sasahara et al., 1991), in particular neurons, and to epithelial cells (Raines and Ross, 1993). The PDGF A chain is more ubiquitously expressed in normal cells as it can be found in smooth muscle cells and fibroblasts throughout the body, in thyroid, muscle, kidney, liver, heart, thymus, and testis as well as the central nervous system (Table 1). The recently cloned PDGF C chain has a unique expression profile that includes the highest level of expression in heart, liver, kidney, pancreas and ovary; smaller amounts of mRNAs in most other tissues, including placenta, skeletal muscle and prostate; and undetectable levels of mRNAs in spleen, colon and peripheral blood leukocytes (Li et al., 2000).
Eliciting and inhibitory stimuli, including exogenous and endogenous modulators Coincident with the tight regulation of PDGF expression in normal cells, a number of eliciting stimuli associated with cellular injury have been identified, as have inhibitory agents that are capable of locally limitating the induction of PDGF (Table 4). The cellular response to PDGF appears to be highly dependent on the local environment independent of the PDGF receptor level. For example, smooth muscle cells (SMCs) cultured on plastic or monomeric type I collagen that normally proliferate in response to PDGF are arrested in G1 and fail to respond to PDGF if plated on polymerized collagen (Koyama et al., 1996). In contrast, fibronectin fibril formation appears to be required for SMC proliferation (Mercurius and Morla, 1998). Thus, the extracellular matrix surrounding the cell can regulate its response to PDGF. We have hypothesized that the matrix surrounding SMCs in the normal media of the vascular wall may be nonpermissive to SMC proliferation (Raines et al., 2000). Polarized secretion, as observed in endothelial cells, which is almost exclusively restricted to the basal compartment (Zerwes and Risau, 1987), can limit the access of surrounding cells to released PDGF. PDGF can also be inhibited by endogenous inhibitors that bind PDGF and inhibit its interaction with its receptors. Plasma binding proteins for PDGF (Raines et al., 1984) include 2-macroglobulin (Raines et al., 1984; Huang et al., 1984), which binds PDGFAB and -BB but not PDGF-AA (Bonner and Osornio-Vargas, 1995). PDGF preferentially binds to native 2-macroglobulin (LaMarre et al., 1991), but the oxidation of 2-macroglobulin, which could
occur in atherosclerotic lesions, decreases the binding of PDGF-BB (Wu et al., 1998) and could thus impair its inhibitory activity. Soluble receptors for PDGF have also been identified in human plasma (Tiesman and Hart, 1993), but little is known about conditions or enzymes responsible for receptor shedding. Other matrix proteins known to bind PDGF include SPARC (Raines et al., 1992), heparin sulfate proteoglycans (Raines and Ross, 1992; Feyzi et al., 1997; Gohring et al., 1998; Lustig et al., 1999), NG2 chondroitin sulfate proteoglycan (Goretzki et al., 1999) and thrombospondin (Hogg et al., 1997). A novel binding protein expressed in a rat neural retina cell line that co-purified with PDGF-AA has been shown to enhance the mitogenic activity of PDGF-AA but decrease the activity of PDGF-BB (Fischer and Schubert, 1996). The specificity of binding and inhibition (Table 5) suggests that more endogenous inhibitors exist for the binding and inhibition of PDGF-AB and -BB than for PDGF-AA. These endogenous binding proteins may help to explain the very rapid clearance (a half-life of less than 2 minutes) of PDGF from the circulation (Bowen-Pope et al., 1984a). Less is known about modulation of the newly identified PDGF-C. However, the fact that it is secreted as an inactive precursor that requires proteolytic removal of the N-terminal CUB domain provides a new level of regulation (Li et al., 2000). Analysis of endogenous and transgenic PDGF-C in heart tissue demonstrates that the protein is processed in vivo. The presence or absence of necessary activating enzymes, therefore, will be critical for induction of PDGF-C activity.
RECEPTOR UTILIZATION As detailed in the PDGF receptor chapter and illustrated in Figure 2, the different dimeric forms of PDGF bind and transmit specific signals via two different receptor subunits that can also form heteroand homodimers with distinct binding patterns (Hart et al., 1988; Heldin et al., 1988; Seifert et al., 1989; Heldin and Westermark, 1999; Li et al., 2000). The capacity of the homo- and heterodimers of PDGF to induce multiple cell functions depends on both the PDGF dimer present and the relative number of the two different PDGF receptor subunits on the responding cells (Seifert et al., 1989; Ferns et al., 1990). PDGF-AA is able to bind only PDGF receptor homodimers, PDGF-AB binds PDGF receptor homodimers and PDGF and receptor heterodimers, while PDGF-BB is able to bind all receptor combinations (Figure 2). The newly identified
766 E. W. Raines Table 4 Eliciting and inhibiting stimuli Stimulus
Target cell/tissue
PDGF induction
Reference
A chain
B chain
ÿ
+
Handa et al., 1998
+
Kirstein et al., 1990
Eliciters AGEs
Epithelial cells Monocytes
-Adrenergic agonists
SMCs
+
ÿ
Majesky et al., 1990b
Blood pressure
Vascular SMCs
+
+
Negoro et al., 1995
IL-1
SMCs and fibroblasts
+
ÿ
Raines et al., 1989; Kawaguchi et al., 1999
TGF
Mammary epithelium
+
ÿ
Bronzert et al., 1990
TGF
SMCs, fibroblasts,
+
ÿ
Battegay et al., 1990
TNF, TGF
Mesangial cells
Cytokines
Osteoclasts +
Astrocytes TNF
Fibroblasts
+
Silver et al., 1989 +
Silberstein et al., 1996
ÿ
Paulsson et al., 1989
+
Keutzer and Sytkowski, 1995
+
Gray et al., 1995
Erythropoietin
Erythroid progenitors
Estrogen
Uterus and vagina
+
Mesangial cells
+
HB-EGF
Endothelial cells
+
PDGF-AA and BB
Bone cell cultures
+
Rydziel et al., 1994
+
Paulsson et al., 1987
Growth factors PDGF, EGF, bFGF, TGF
PDGF Hypoxia
Brain, neonatal rat
Lipopolysaccharide
Endothelial cells, human pulmonary
PDGF
Fibroblasts
+
Osteoblasts
+
Pressure
Mesangial cells
ÿ
Shear stress
Endothelial cells
Step and impulse
ÿ
Silver et al., 1989 ÿ
Gay and Winkles, 1990
+
Ohno et al., 1999
+
Albelda et al., 1989
+
Paulsson et al., 1987 Rydziel et al., 1994
+
Kato et al., 1999
+
Resnick et al., 1993; Khachigian et al., 1995
+
Bao et al., 1999 Wilson et al., 1998
Strain
SMCs
+
Thrombin
Mesangial cells
+
SMCs
+
ACE inhibitors
Endothelial cells
+
+
Yamaguchi et al., 1994
+
+
Wong et al., 1997
Calcium channel blockers
Endothelial cells
+
+
Yamaguchi et al., 1994
Carbon monoxide
Endothelial cells
+
Morita and Kourembanas, 1995
Glucocorticoid
SMCs
+
Nakano et al., 1993
Oxidized LDL
Monocytes
+
Malden et al., 1991
+
Shultz et al., 1989 Okazaki et al., 1992
Inhibitors
+
PDGF 767 Table 5 Binding proteins for different dimeric forms of PDGF Binding protein
PDGF-AA long
PDGF-AA short
PDGF-BB
PDGF-AB
Inhibition or enhancement
2-Macroglobulin
ND
ÿ
+
+
ÿ
Heparan sulfate proteoglycan
+
ÿ
+
ND
ÿ
NG2 proteoglycan
ND
+
ÿ
ND
ÿ
PDGF-associated protein (PAP)
ND
+
PAP
ÿ +
ND
+
SPARC
ÿ
ÿ
+
+
ÿ
Thrombospondin
ND
ND
+
ND
+
Figure 2 The binding of PDGF to its receptor subunit pairs depends on the chain composition of the PDGF dimer. Two types of PDGF receptor subunits exist: PDGF receptor that can bind PDGF A- and Bchains, and PDGF -receptor which only binds PDGF B-chain. Thus, PDGF-BB can bind to all three subunit pairs of PDGF receptors ± receptors, -receptors, and -receptors; PDGF-AB can bind two subunit receptor pair ± -receptors and receptors; while PDGF-AA is only able to bind only to -receptors.
PDGF-CC only binds PDGF receptor homodimers (Li et al., 2000).
IN VITRO ACTIVITIES
In Vitro findings The in vitro activities of PDGF have been extensively characterized, especially for mesenchymal cells, whose stimulation served as the basis for its purification.
PDGF is a potent stimulant of the proliferation of mesenchymal cells and stem cells of multiple lineages, but a more limited stimulant of endothelial and epithelial cells. It is also a potent stimulant of the proliferation of vascular smooth muscle cells (Ross et al., 1974), kidney mesangial cells (Shultz et al., 1988), dermal fibroblasts (Heldin et al., 1981b), glial cells (Heldin et al., 1981b), testicular peritubular myoid cells (Gnessi et al., 1992), muscle cells (Jin et al., 1991), osteoblasts (Centrella et al., 1991), and chondrocytes (Raines et al., 1990; Kieswetter et al., 1997). In many cases, such as osteoblasts and chondrocytes, the PDGF stimulation of proliferation prevents differentiated function (Hock and Canalis, 1994; Kieswetter et al., 1997). Within the nervous system, PDGF released by type 1 astrocytes stimulates O-2A progenitor cells in developing rat optic nerve, which gives rise to oligodendrocytes and type 2 astrocytes (Noble et al., 1988; Raff et al., 1988). Schwann cells also proliferate in response to PDGF in the presence of forskolin or other agents that raise the intracellular cyclic AMP level (Davis and Stroobant, 1990). Some of the growth-promoting activity of PDGF appears to be indirect and mediated through the induction of other growth stimulants specific for the cell types (Raines et al., 1990). Examples include the PDGF enhancement of erythropoietic progenitor cell proliferation (Dainiak et al., 1983; Delwiche et al., 1985), the promotion of T cell proliferation (Daynes et al., 1991), and the stimulation of primitive hematopoietic precursors (Yan et al., 1993). Although PDGF has been reported to be angiogenic in vivo (Risau et al., 1992), it appears primarily to stimulate angiogenesis by locally stimulating other potent angiogenic stimulants, such as VEGF (D. Wang et al., 1999) and matrix, such as type I collagen (Sato et al., 1993). However, PDGF-BB has been shown to induce functional vascular anastomoses in vivo (Brown et al., 1995), regardless of a direct or indirect mechanism.
768 E. W. Raines The first report of the PDGF stimulation of an epithelial tissue was its stimulation of growth of the lens in an organ culture system (Brewitt and Clark, 1988). A proliferative and migratory response to PDGF-BB has been demonstrated for rabbit gastric epithelial cells (Watanabe et al., 1996).
products. The induction of other cytokines following PDGF stimulation, including its own induction (Paulsson et al., 1987), may serve as a means to amplify its proliferative effects.
PDGF as a Survival Factor
The chemotaxis of arterial smooth muscle cells in a Boyden chamber assay with collagen-coated filters is stimulated by all dimeric forms of PDGF and is dependent upon the relative expression levels of the different PDGF receptors (Ferns et al., 1990). PDGF also stimulates the migration of fibroblasts (Seppa et al., 1982), human keratinocytes (Andresen and Ehlers, 1998), human retinal pigment epithelial cells (Campochiaro and Glaser, 1985), and neuroepithelial stem cells (Forsberg-Nilsson et al., 1998). PDGF as a potential chemoattractant for neutrophils and monocytes (Deuel et al., 1982; Williams et al., 1983) is more controversial (Graves et al., 1989), which may be because of a requirement for activation of the monocytes and neutrophils by cytokines or lymphocytes to induce PDGF responsiveness (Shure et al., 1992).
Since PDGF promotes the proliferation of connective tissue-synthesizing cells, the ability of PDGF to modify the matrix was examined shortly after its purification. PDGF composed of a mixture of the different isoforms was shown both to promote matrix synthesis via the stimulation of collagen and thrombospondin synthesis, and to promote matrix breakdown via the promotion of collagenase activity (Raines et al., 1990). It has subsequently been shown that a number of other extracellular matrix constituents are regulated by PDGF, including the stimulation of hyaluronan synthesis (P. Heldin et al., 1989), of versican chondroitin sulfate proteoglycans (SchoÈnherr et al., 1991), of the small chondroitin sulfate/dermatan sulfate proteoglycans biglycan and decorin (SchoÈnherr et al., 1993), of lysyl oxidase, which may modulate matrix organization (Green et al., 1995), of the alternative splicing of fibronectin, which alters its cell-binding properties (McKay et al., 1994), of alternatively spliced forms of tenascin (LaFleur et al., 1994), and of both plasminogen activators (Pfeilschifter et al., 1992; Kenagy and Clowes, 1995) and the receptor for urokinase-type plasminogen activator receptor (Reuning and Bang, 1992). It is likely that the induction of these matrix constituents may be critical for many of the effects of PDGF, as has been demonstrated for smooth muscle cell migration and proliferation, which is dependent on urokinase and tissue-type plasminogen activator (Herbert et al., 1997).
Induction of `Early Genes'
PDGF Modulation of Specific Cellular Functions
The PDGF stimulation of responsive cells regulates the expression of a discrete group of low-abundance and labile gene products (Rollins and Stiles, 1988). At least three of these immediate early genes ± c-fos, cmyc, and Egr-1 ± appear to act as intracellular mediators of the PDGF mitogenic response (Kelly et al., 1983; Cochran et al., 1984; Greenberg and Ziff, 1984; Santiago et al., 1999). In the case of neuroepithelial cells from the developing cortex, the immediate early gene response is sufficient to specify the neuronal fate (Williams et al., 1997). JE and KC are two other `early genes' induced by PDGF (Rollins et al., 1988) that encode monocyte chemotactic protein 1 (Rollins, 1991) and melanocyte growth factor (Oquendo et al., 1989) respectively, which are proinflammatory gene
The effects of PDGF on specific cellular functions are dependent on the cell type. In oligodendrocytes, PDGF relieves the astrocyte inhibition of myelin basic protein mRNA transport into cell processes and cytoplasmic channels that infiltrate the myelin sheath, important for myelination (Amur-Umarjee et al., 1997); in 3T3-L1 adipocytes, PDGF promotes glucose uptake by translocation of the glucose transporter (L. Wang et al., 1999); cationic amino acid transporter gene expression is stimulated in smooth muscle cells and supports proliferation (Durante et al., 1996); PDGF stimulates mesangial cell ion channels that would promote sustained extracellular calcium entry in human skin fibroblasts (Ling et al., 1995), and modulates junctional cell-to-cell communication
During development, cell turnover is modulated by both proliferation and apoptosis. In the developing rat optic nerve, about 50% of the oligodendrocytes normally die as a result of competition for a limited number of survival signals, one of which is PDGF (Barres et al., 1992). PDGF-AA and -BB are also survival factors for neurons from ventral mesencephalon from rat and human embryos (Nikkhah et al., 1993), developing Schwann cells (Meier et al., 1999), and a rat hippocampal cell line (Kwon, 1997). PDGF Stimulates Directed Cellular Migration
Modification of Cell Matrix Constituents by PDGF
PDGF 769 (Maldonado et al., 1988; Hossain et al., 1999); both PDGF-AA and -BB induce the low-density lipoprotein receptor gene (Rechtoris and Mazzone, 1995); PDGF inhibits human natural killer cell activity (Gersuk et al., 1991); and it enhances the expression of -amino-e-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA)-type glutamate receptor 1 in endocortical neurons through the Src family kinase Fyn (Narisawa-Saito et al., 1999); in blood monocytes, PDGF-BB induces the expression of the tissue factor that may promote inflammation (Ernofsson and Siegbahn, 1996); plasminogen activator inhibitor 1 stimulation by PDGF in smooth muscle cells may regulate the fibrinolytic system (Reilly and McFall, 1991); and in human trabecular cells, PDGF facilitates phagocytosis (Tamura and Iwamoto, 1989).
Regulatory molecules: Inhibitors and enhancers As discussed above, the response of cells expressing PDGF receptors is highly dependent on the extracellular matrix environment and on the level of endogenous binding proteins for PDGF (see Table 5), which bind and inhibit PDGF binding to responsive cells. Some of the binding proteins, such as the PDGF-associated protein (Fischer and Schubert, 1996) and thrombospondin (Hogg et al., 1997), do not interfere with receptor binding and appear to enhance PDGF activity.
Bioassays used The principal bioassays used to evaluate PDGF activity are the PDGF radioreceptor assay (Bowen-Pope and Ross, 1985) and the stimulation of connective tissue proliferation in the presence and absence of PDGF neutralizing antibodies, as, for example, reported for brain-derived PDGF (Sasahara et al., 1991).
IN VIVO BIOLOGICAL ACTIVITIES OF LIGANDS IN ANIMAL MODELS
Normal physiological roles
development (reviewed in Betsholtz, 1995; Ataliotis and Mercola, 1997; Lindahl and Betsholtz, 1998; Heldin and Westermark, 1999; Li et al., 2000). A common theme observed in both mammalian development as well as in Xenopus (Jones et al., 1993; Ghil and Chung, 1999) and sea urchin (Ramachandran et al., 1997) development is that ectodermally produced PDGF may act on mesoderm during gastrulation. Mesenchymal cells throughout the embryo and within particular developing organs express PDGF receptors, while PDGF A, B, and C chains are often secreted by adjacent epithelial cells or endothelial cells. Pericytes and mesangial cells failed to develop in PDGF B chain-deficient embryos (Lindahl et al., 1997a; LeveÂen et al., 1994), whereas alveolar smooth muscle cells failed to develop in PDGF A chaindeficient embryos (BostroÈm et al., 1996; Lindahl et al., 1999). PDGF B chain also acts locally to contribute to the development of the labyrinthine layer of the fetal placenta and the formation of a proper nutrient± waste exchange system during fetal development (Ohlsson et al., 1999). A reduction in number of both pericytes and trophoblasts is observed in embryos deficient in either PDGF B chain or the PDGF receptor. Although PDGF-CC shares the same receptor with PDGF-AA, it appears to have a distinct expression profile and, thus, different roles (Li et al., 2000). This was predicted from PDGF receptor-null animals that show certain defects, such as cleft face and spina bifida, that are not seen in homozygous deletions of the genes for PDGF-A, PDGF-B, or both. Mouse embryos lacking PDGF receptor show an extensive loss of the cortical mesenchyme adjacent to sites of PDGF-C-expressing cells in early aggregates of metanephric mesenchyme undergoing epithelial conversion in the kidney (Li et al., 2000). Thus, it has been proposed that PDGF-C has an essential function in the development of the kidney mesenchyme. Patterns of PDGF ligand and receptor expression in the developing palate (Qiu and Ferguson, 1995), kidney (Alpers et al., 1992, 1995; Seifert et al., 1998), teeth (Chai et al., 1998), testis (Gnessi et al., 1995), myoblasts (Jin et al., 1990), retina (Mudhar et al., 1993; Frutigger et al., 1996), lens (Potts et al., 1994), rat optic nerve (Richardson et al., 1988; Barres et al., 1992), and spinal cord (Pringle and Richardson, 1993; Hall et al., 1996) are consistent with a role for PDGF in embryogenesis via both autocrine and paracrine signaling.
In the Developing Embryo
Nervous System
A large number of studies have demonstrated distinct roles for PDGF A, B, and C chain during
The importance of PDGF in the proliferation and migration of mesenchymal cells of the brain, glial
770 E. W. Raines cells, served as one of the bases for its purification (Westermark and Wasteson, 1976). However, its potential as a broad modulator of the nervous system was not suspected until a survey of normal tissues demonstrated a very significant expression of the PDGF B chain in neurons, principal dendrites, some axons, and probable terminals throughout the brain, as well as in the dorsal horn of the spinal cord and the posterior pituitary (Sasahara et al., 1991). Similarly, the expression of PDGF A chain mRNA is observed to have a similar distribution in adults (Sasahara et al., 1991; Yeh et al., 1991). These observations have led to a detailed examination of the responses of specific cells within the nervous system to PDGF, both in vitro and in vivo, and have led to the realization that PDGF is a central neurotrophic factor (Valenzuela et al., 1997). Although the significance of its actions are not totally understood, ligand-gated ion channels, both type A aminobutyric acid (GABAA) and N-methyl-D-aspartate (NMDA) receptors on neurons, are modulated by PDGF (Valenzuela et al., 1995, 1996) and may regulate synaptic plasticity in the brain (Valenzuela et al., 1997). More recently, PDGF has also been shown to increase the number of -amino-3-hydroxy5-methyl-4-isoxazolepropionic acid (AMPA)-type glutamate receptors (Narisawa-Saito et al., 1999), which may also contribute to the regulation of neuronal plasticity. PDGF has also been implicated in the protection of neurons. The administration of PDGF-BB prevents spinal motor neuron death in neonatal rats following sciatic nerve resection (Iwasaki et al., 1997) and delays neuronal death in CA1 pyramidal neurons after forebrain ischemia (Iihara et al., 1997; Kawabe et al., 1997). This is in agreement with the ability of PDGF-BB to enhance neuron survival following the intraocular transplantation of E14 ventral mesencephalon to sympathetically denervated host eyes (Giacobini et al., 1993), to reduce the loss of dopaminergic neurons in cultures from rat and human embryonic ventral mesencephalon (Nikkhah et al., 1993), to increase the survival of GABAergic neurons in a primary culture of newborn rat cerebellum (Smits et al., 1991), and to stimulate the neuronal cell survival factor, brain-derived neurotrophic factor, in a rat hippocampal stem cell line (Kwon, 1997). A potential role for PDGF in the modulation of neuronal energy metabolism has been suggested by the ability of PDGF to induce phosphorylation of the delta subunit of the F1F0 ATPase in cultured mouse cortical neurons (Zhang et al., 1995). Another role for PDGF in the nervous system was implied by our observations that PDGF inhibits the feeding response in the coelenterate Hydra (Hanai et al., 1987), one of
the earliest animals known to have a simple nerve net. The ability of PDGF to alter the feeding response has been confirmed in mammals by the intracerebroventricular injection of PDGF (Plata-SalamaÂn, 1988; Sasaki et al., 1991). In mammals, it appears that regulation of the feeding response by PDGF may be specific to particular phases. Normal Wound Repair Normal wound healing requires the concerted actions of a number of cell types. The earliest cells that appear in wounds are leukocytes, principally neutrophils together with monocytes, which become the dominant cell within a day or two following wounding. Subsequently, the ingress and proliferation of capillaries, fibroblasts, and smooth muscle cells ultimately leads to wound resolution (Ross, 1968). In addition to stimulating the migration and proliferation of particularly fibroblasts into the wound site, PDGF can modulate local gene expression (Rollins and Stiles, 1988; Raines et al., 1990), including that of cellular matrix constituents (Raines et al., 1990) and coagulation factors (Ernofsson and Siegbahn, 1996; Edelberg et al., 1998), and can regulate interstitial fluid pressure by altering tension on the extracellular matrix structures of fibroblasts and other connective tissue cells (Rodt et al., 1996; Heuchel et al., 1999). PDGF can be made by all of the cell types involved in wound repair, with the exception of the neutrophils, and the expression of PDGF depends on the local stimulants. Although circulating monocytes express a low or undetectable level of PDGF (Ross et al., 1990), infiltrating macrophages express PDGFBB/AB (Reuterdahl et al., 1993). Epithelial cells are also a major source of PDGF (see Table 1) and may thus be a significant source of PDGF following wounding of the epidermis (Ansel et al., 1993). PDGF expression is also observed in peripheral nerve fibers in wounds (Reuterdahl et al., 1993). Numerous studies have demonstrated the presence of PDGF in wound fluid in partial thickness wounds (Vogt et al., 1998), following mastectomy (Dvonch et al., 1992), during fracture repair (Andrew et al., 1995; Fujii et al., 1999), in tear fluid following photorefractive keratectomy (Vesaluoma et al., 1997), and in blisters of partial skin thickness burns (Ono et al., 1995). It has also been noted that, in addition to the presence of PDGF-BB and -AB, a markedly upregulated level of the long form of PDGF-AA, which preferentially binds the extracellular matrix, is observed in acute, surgically created wounds (Pierce et al., 1995). Furthermore, a reduced expression of PDGF and its receptors has been noted during impaired wound healing (Pierce et al., 1995; Beer et al., 1997).
PDGF 771 The expression of PDGF during the repair of injury is not limited to external wounds as it is also observed following arterial injury (Majesky et al., 1990a), periodontal wound healing (Lynch et al., 1991), bone fracture (Andrew et al., 1995; Fujii et al., 1999), nerve injury (Mekada et al., 1998; Ohno et al., 1999), kidney injury (Iida et al., 1991), and hypoxic lung injury (Katayose et al., 1993). Because of the critical problem of defective wound repair in patients who are bedridden for long periods of time, or who may have other chronic diseases, such as diabetes, that interfere with the healing process, several investigators have examined the effects of PDGF on granulation tissue formation in models of impaired wound healing (Grotendorst et al., 1985; Greenhalgh et al., 1990; Pierce et al., 1991, 1995; Liechty et al., 1999). PDGF significantly augments the time-dependent influx of inflammatory cells and fibroblasts, and accelerates provisional extracellular matrix deposition and subsequent collagen formation. In a tracheal wounding model, administration of PDGF-BB in a collagen±fibrin composite accelerated wound repair (Koempel et al., 1998). Although most studies utilized recombinant PDGF-BB (which binds and activates all forms of the PDGF receptor), the overexpression of PDGF-AA in genetically modified human keratinocytes has also been shown to enhance the performance of a skin graft (Eming et al., 1995, 1998). A comparison of the behavior of wild-type and null PDGF receptor cells within individual chimeric mice has revealed that the PDGF B chain is important for fibroblast and endothelial cell recruitment into sponge implants (Crosby et al., 1999). PDGF as a Possible Physiologic Mediator of Smooth Muscle Cell Function Changes in PDGF A chain expression in both vascular smooth muscle (Majesky et al., 1990b) and uterine smooth muscle (Boehm et al., 1990; Mendoza et al., 1990; Gray et al., 1995) associated with changes in their function suggest that PDGF may be a physiologic mediator of these functions (Raines and Ross, 1993). In the rat aorta, -adrenergic stimulation dose-dependently stimulates PDGF A chain gene expression (Majesky et al., 1990b). The induction of PDGF A chain was specific to 1adrenergic receptors as angiotensin II and endothelin, despite an increase in blood pressure, had little or no effect on the PDGF A chain level. Adrenergic stimulation of the PDGF A chain was coincident with increases in growth-related genes, but not DNA synthesis, which were associated with the trophic effects of sympathetic nerves and catecholamines on arterial smooth muscle cell mass and protein synthesis.
In a one-kidney, one-clip hypertensive rat model of nonrenin-dependent hypertension, PDGF A chain expression correlates with blood pressure and increased wall area (Dobrian et al., 1999). High blood pressure in spontaneously hypertensive rats (SHRs) is also associated with an increased level of the PDGF A chain (Negoro et al., 1995), and the administration of an antisense oligodeoxynucleotide to the PDGF A chain in SHRs did not alter the systolic blood pressure but markedly reduced the incorporation of [H3]thymidine into aortic DNA and suppressed elevated DNA content (Fukuda et al., 1997a). Interestingly, the long form of the PDGF A chain has been shown to be elevated in SHRs (Fukuda et al., 1997b). A comparison of human myometrium from normal uteri and gestational uteri demonstrated a 10- to 15fold increase in PDGF A chain expression at term and was associated with uterine expansion (Mendoza et al., 1990). An analysis of cell-specific RNA expression revealed that estrogen treatment decreases the level of PDGF and its receptors in both the uterus and vagina of the mouse (Gray et al., 1995). PDGF is also expressed in smooth muscle cell-like Leydig cells in the testis, it stimulates proliferation and matrix production (Gnessi et al., 1993), and its secretion is dose-dependently increased by the trophic hormone human chorionic gonadotropin (Gnessi et al., 1992). Thus, the hormonal regulation of uterine and testicular PDGF may mediate smooth muscle cell function. Other Possible Vascular Regulation by PDGF Although PDGF significantly modulates the functions of all smooth muscle cells and pericytes, as well as of some endothelial cells, PDGF does not appear to be absolutely required for particularly large vessel formation as a normal vascular network is observed in knockouts of both PDGF ligands and their receptors (LeveÂen et al., 1994; Soriano, 1994, 1997; BostroÈm et al., 1996). However, the PDGF B chain is particularly critical for the recruitment of vascular SMCs and pericytes, as shown by an analysis of PDGF B chain and PDGF receptor-null mice (Lindahl et al., 1997a; HellstroÈm et al., 1999), and PDGF receptor-null cells in mouse chimeras (Crosby et al., 1998). Although a role in adult animals has not been tested, it has been proposed that the PDGF B chain may play a similarly important role in these, for example in the venous wall that has only a rudimentary SMC coating before birth (HellstroÈm et al., 1999). The PDGF B chain is important for endothelial cell recruitment into sponge implants, as recently demonstrated by a comparison of the behavior of wild-type and null PDGF receptor cells within individual chimeric mice (Crosby et al., 1999).
772 E. W. Raines The role of PDGF in the regulation of vascular tone is less clear. In strips of rat aorta in vitro, a mixture of PDGF isoforms from platelets induces dose-dependent contraction (Berk et al., 1986). In contrast, rat intracerebral arterioles isolated from brain parenchyma did not respond to any of the dimeric forms of PDGF (Bassett et al., 1988), suggesting vascular bed-specific responses that may be dependent on PDGF receptor expression. PDGF-BB has also been reported to stimulate the nitric oxidemediated relaxation of rat aortic rings (Cunningham et al., 1992; Takase et al., 1999), which is consistent with data in which only PDGF-BB infusion into anesthetized rats lowered blood pressure (Ikeda et al., 1997). These differences are likely to reflect complex interactions between the endothelium and SMCs that may vary in different vascular beds. PDGF receptor expression on endothelium appears to be limited to capillary endothelial cells (Bar et al., 1989; Smits et al., 1989; Beitz et al., 1991; P. Heldin et al., 1991), and local PDGF expression adjacent to capillaries may control their function. Endothelial cells of different vascular beds synthesize discrete gene products associated with organ-specific tasks (Sawdey and Loskutoff, 1991; Weiler-Guettler et al., 1996; Aird et al., 1997). In the mouse heart, the myocyte modulation of PDGF expression by microvascular endothelial cells regulates the expression of hemostatic and angiogenic gene products, including von Willebrand factor, VEGF and its receptor, flk-1 (Edelberg et al., 1998). The localized expression of PDGF in brain, heart, and muscle, by astrocytes, cardiac myocytes, and skeletal myocytes respectively, also regulates endothelial nitric oxide synthase (Guillot et al., 1999). Thus, the actions of PDGF on genes involved in the regulation of angiogenesis and vascular tone may help to explain the variability of its purported actions in these two processes. Potential Role of PDGF in the Expansion of Stem Cell Populations During development, PDGF appears to play a major role in the expansion of certain stem cell populations, such as O2-A progenitor cells (Raff et al., 1988). Similarly in adults, PDGF may be important for the expansion of adult O-2A progenitor cells (Wolswijk and Noble, 1992) and of a pool of multipotent cell precursors that resides within the cerebral cortex (Marmur et al., 1998). PDGF also enhances the proliferation of myoblasts (Jin et al., 1991; YablonkaReuveni and Seifert, 1993) and may be important in the transition of adult satellite cells to proliferating myoblasts (Yablonka-Reuveni, 1995). This possibility could be tested in mice created from chimeric
blastocysts composed of a mixture of wild-type and PDGF receptor-null cells, which have demonstrated an 8-fold reduction in the participation of PDGF receptor-null cells in all muscle lineages (Crosby et al., 1998). PDGF may also influence circulating stem cell populations. Primitive multipotent hematopoietic precursors (pre-CFC multi) are stimulated by PDGF (Yan et al., 1993), as are erythroid precursors (Dainiak et al., 1983; Delwiche et al., 1985), both via indirect mechanisms in which they stimulate cytokine secretion from other cells. PDGF also stimulates megakaryocyte precursor growth (Yang et al., 1995), the precursors of platelets. Although a direct link to precursor growth has not been demonstrated, early B lineage precursor cells express PDGF receptors (Tsai et al., 1994; Trink et al., 1995) and may thus contribute to early B progenitor expansion.
Species differences The PDGF molecules appear to be highly conserved. An analysis of Xenopus laevis genomic DNA using human PDGF A and B chain probes demonstrated homologous species with 73% identity over the A chain coding region, including evidence for both the long and short forms of the A chain, and 53% and 65% sequence identity in the 50 and 30 UTRs respectively (Mercola et al., 1988). Phylogenetic analyses of clotted whole blood have demonstrated that PDGFBB is the predominant isoform in sera from mouse, rat, pig, cow, sheep, dog, and chicken, but not primates (Bowen-Pope et al., 1989), and that blood from all chordates contains PDGF while sera from tunicates to the arthropod line of development were negative (Singh et al., 1982). However, functional homologs of PDGF and its receptor appear to be present in even more primitive animals as the protozoan Tetrahymena responds chemotactically to PDGF and the small freshwater coelenterates Hydra release a PDGF homolog on wounding that inhibits their feeding response and is blocked by an antibody to human PDGF (Hanai et al., 1987; Raines et al., 1990). PDGF protein has also been detected in invertebrate and vertebrate immunocytes (Franchini et al., 1996).
Knockout mouse phenotypes A deletion of either the PDGF A or B chain is embryonic lethal (LeveÂen et al., 1994; BostroÈm et al., 1996). The phenotype of the knockout embryos has pointed to the particular importance of the PDGF B chain for the development of kidney glomeruli
PDGF 773 (LeveÂen et al., 1994; Lindahl et al., 1998) and suggests that although the mesangial cell lineage is independent of the PDGF B chain, it is required for the recruitment of PDGFR -positive progenitors surrounding the developing glomerular afferent and efferent arterioles during angiogenic formation of the glomerular capillary tuft (Lindahl et al., 1998, 1999). PDGF B chain knockouts have also demonstrated the importance of PDGF B chain in the recruitment of vascular smooth muscle cells and pericytes during embryonic blood vessel formation (Lindahl et al., 1997a; HellstroÈm et al., 1999). The recruitment of alveolar smooth muscle cells, important for elastin fibre formation, is also a major defect in PDGF A chain-deficient embryos (BostroÈm et al., 1996; Lindahl et al., 1997b). Expression patterns and mutant phenotypes suggest that epidermal cell PDGF A chain expression is important for dermal mesenchymal proliferation and may contribute to the formation of dermal papillae, mesenchymal sheaths and dermal fibroblasts (Karlsson et al., 1999). As predicted from the dependence of oligodendrocyte development on astrocyte production of the PDGF A chain (Noble et al., 1988; Raff et al., 1988), embryos lacking the PDGF A chain have defective oligodendrocyte development and severe hypomyelination (Fruttiger et al., 1999). Many of the phenotypes in PDGF A and B chain knockouts are also observed in PDGF receptor knockouts (Soriano, 1994, 1997). The presence of additional defects in PDGF receptornull embryos as compared with homozygous deletions of the genes for PDGF-A, PDGF-B, or both, appears to be explained by the identification of the PDGF-C gene (Li et al., 2000).
Transgenic overexpression As would be predicted from the potent activities of PDGF on mesenchymal cells and knockout studies of both PDGF chains, the overexpression of PDGF particularly stimulates connective tissue cell accumulation. Overexpression of the PDGF B chain in hematopoietic cells in mice induces a lethal myeloproliferative syndrome (Yan et al., 1994); overexpression of the PDGF A chain under the control of the lens-specific A crystallin promoter result in cataracts and retinal astrocyte hyperplasia (Reneker and Overbeek, 1996a, 1996b); keratinocyte-produced PDGF-AA promotes epidermal±dermal interactions and promotes the growth and vascularization of dermal tissue (Eming et al., 1995), as well as improving graft performance (Eming et al., 1998); while PDGF-BB accelerates gastric epithelial cell restoration (Watanabe et al., 1996) and corrects ischemic
impaired wound healing (Liechty et al., 1999). Transgenic overexpression of PDGF-C in mice using the -myosin heavy chain promoter induced strong proliferation of myocardial interstitial cells, such as cardiac fibroblasts (Li et al., 2000).
Interactions with cytokine network The PDGF stimulation of responsive cells induces the expression of a number of cytokines as part of the `early gene' response, as detailed above. A number of cytokines can also regulate the expression of PDGF (see Table 4 and above).
Endogenous inhibitors and enhancers As discussed above, the response of cells expressing PDGF receptors is highly dependent upon the extracellular matrix environment and on the level of endogenous binding proteins for PDGF (see Table 5), which bind and inhibit PDGF binding to responsive cells. Some of the binding proteins, such as the PDGF-associated protein (Fischer and Schubert, 1996) and thrombospondin (Hogg et al., 1997), do not interfere with receptor binding and appear to enhance PDGF activity.
PATHOPHYSIOLOGICAL ROLES IN NORMAL HUMANS AND DISEASE STATES AND DIAGNOSTIC UTILITY
Normal levels and effects As we have already discussed, although PDGF may be expressed by a multiplicity of normal cells and tissues (see Table 1), its expression in normal cells is highly regulated. PDGF has a number of effects in developing embryos, in the nervous system, in normal wound repair, as a physiological mediator of SMC function, in vascular regulation, and in the expansion of stem cell populations as discussed above.
Role in experiments of nature and disease states The expression of PDGF and its receptors is increased in a number of disease states (see Table 2), including neoplasias of multiple cell origins (see
774 E. W. Raines Table 3). Whereas normal tissue repair results in simple scar formation, prolonged inflammation may result in the excessive accumulation of connective tissue cells and extensive matrix deposition, with potentially pathological consequences. As discussed below, for several of the disease states listed in Table 2 above, a common feature of those diseases associated with an increased expression of PDGF and its receptors is the recruitment and stimulation of the local SMC population. Atherosclerosis and Restenosis Atherosclerosis remains the leading cause of death despite changes in lifestyle and effective pharmacologic regimens to lower cholesterol level, one of the principal risk factors for atherosclerosis (Breslow, 1997). A common feature that atherosclerosis shares with a number of other disease processes is that it is an inflammatory disease (Ross, 1999). Although a number of growth factors and cytokines in addition to PDGF are increased in the lesions of atherosclerosis (Raines and Ross, 1996), recent evidence in a nonhuman primate model of restenosis (Giese et al., 1999) provides support for previous studies in other animal models (Raines and Ross, 1993) that the blockade of PDGF receptors may reduce neointimal lesion formation. PDGF expression is increased in circulating mononuclear cells of hypercholesterolemic patients (Billett et al., 1996), and PDGF B chain expression is observed in macrophages in lesions at all stages of atherosclerosis (Ross et al., 1990). In advanced lesions in nonhuman primates, PDGF A and B chain expression is observed in both macrophages and SMCs (Evanko et al., 1998). Expression of the PDGF receptor is increased following angioplasty in human (Tanizawa et al., 1996) and nonhuman primate (Giese et al., 1999), and blockade of the PDGF receptor inhibited SMC neointimal accumulation by 37% following angioplasty (Giese et al., 1999). PDGF and its receptors are also increased in the macrophage-rich matrix surrounding polytetrafluoroethylene vascular grafts in baboons (Kraiss et al., 1993) as well as in vein grafts in pigs (Francis et al., 1994). An acute reduction in blood flow and shear stress induces PDGF A chain in the baboon prosthetic grafts (Kraiss et al., 1996), while stenting in the pig arteriovenous bypass graft model was associated with a reduced intimal formation and reduced expression of PDGF (Mehta et al., 1998). Glomerulonephritis As observed in patients with hypercholesterolemia (Billett et al., 1996), PDGF B chain expression is
increased in peripheral blood mononuclear cells in IgA nephropathy and has been shown to be positively correlated with protein secretion (Nakamura et al., 1992). A further analysis of renal biopsy specimens demonstrated that glomerular proliferation as well as clinical measures of disease severity were greater in patients positive for PDGF A and B chain (Taniguchi et al., 1996). Similarly, a marked increase in the expression of the PDGF receptor and a more modest increase in PDGF receptor correlated with the grade of histologic lesion in IgA nephropathy (Gesualdo et al., 1994). Thus, although PDGF is required for glomerular development, its sustained overexpression in human disease and a number of animal models of glomerulonephritis suggests that PDGF may be particularly involved in mesangial cell proliferation and connective tissue deposition (Raines and Ross, 1993; Abboud, 1995; Heldin and Westermark, 1999). The infusion of PDGF-BB, but not PDGF-AA, into normal rats stimulated SMC interstitial proliferation and tubulointerstitial fibrosis (Tang et al., 1996). In the rat Thy-1 model of glomerulonephritis, the infusion of PDGF-BB increased glomerular cell proliferation 32-fold, as well as the glomerular deposition of type IV collagen, laminin, and fibronectin (Floege et al., 1993). Antagonism of PDGF in the same model, with either anti-PDGF antibodies (Johnson et al., 1992) or a PDGF B chain aptamer (Floege et al., 1999), significantly reduced the proliferation of mesangial cells and the glomerular deposition of type IV collagen and fibronectin. Proliferative Retinal Disease Proliferative retinopathy is a common complication of diabetes, a disease in which the increased release of PDGF from platelets has been suggested (Guillausseau et al., 1989). An analysis of the human proliferative retinal membranes of individuals with diabetes demonstrated a significantly increased PDGF A chain level in retinal pigment epithelial cells as well as PDGF B chain level in retinal membranes, while primarily vascular cells were positive for both receptors and, principally, PDGF receptor expressed on retinal pigment epithelial cells (Robbins et al., 1994). A similar distribution was observed in proliferative vitreoretinopathy patients. In a rabbit model of proliferative vitreoretinopathy, the injection of embryo fibroblasts derived from PDGF receptor knockout embryos into the eyes of rabbits that had previously undergone gas vitrectomy failed to induce disease (Andrews et al., 1999). The injection of cells expressing PDGF receptor, but not receptor, replaced the ability to induce proliferative vitreoretinopathy.
PDGF 775 Pulmonary Fibrosis PDGF A and B chain expression is increased in alveolar macrophages of individuals with idiopathic pulmonary fibrosis (Nagaoka et al., 1990) and histiocytosis X (Uebelhoer et al., 1995). The intratracheal injection of PDGF-BB into rats (Yi et al., 1996) and the overexpression of PDGF-BB under the control of the lung-specific surfactant protein C in transgenic mice (Hoyle et al., 1999) stimulate pulmonary mesenchymal and epithelial cell proliferation, collagen deposition and lung fibrosis. Data from the vanadium pentoxide-induced rat injury model demonstrate that expression of the PDGF receptor, but not the receptor, is increased in interstitial SMCs in the lung and precedes the development of the fibroproliferative lung lesion (Bonner et al., 1998). PDGF receptor-inducing activity was present in alveolar macrophages following injury, this inducing activity being blocked by the IL-1 receptor antagonist. Both IL-1 and TNF are increased following lung injury, and mice lacking TNF fail to develop fibroproliferative lesions after asbestos exposure; in addition, their expression of PDGF A chain is reduced (Liu et al., 1998). More specific data supporting a role for PDGF are provided by use of AG1296, the tyrosine kinase inhibitor for the PDGF receptor, in the vanadium pentoxide model of lung injury in which AG1296 reduced lung cell proliferation and was more than 90% effective in preventing the increase in hydroxyproline used as a measure of collagen deposition (Rice et al., 1999). Given the critical role of the PDGF receptor in lung alveogenesis (BostroÈm et al., 1996; Lindahl et al., 1997b) and the induction of the PDGF receptor (Bonner et al., 1998), it would be interesting to test the efficacy of a specific PDGF receptor antagonist in the lung fibrosis. Transplant Rejection PDGF A and B chain levels are increased in human renal rejection (Alpers et al., 1996), as are those of the PDGF and receptors (Fellstrom et al., 1989; Floege et al., 1998). Consistent with the T cell induction of PDGF expression in endothelial cells (Shaddy et al., 1992), PDGF A chain expression is increased in endothelial cells and intimal and medial SMCs in renal vascular rejection, with less prominent PDGF B chain expression in intimal and medial SMCs (Alpers et al., 1996). A strong focal expression of PDGF receptor expression was observed in the intimal SMCs of vessels exhibiting signs of arteriosclerosis in areas adjacent to increased PDGF A chain levels (Floege et al., 1998). Given the prominent expression of the PDGF receptor and PDGF A chain, it would be
interesting to test a PDGF receptor-specific antagonist. Chronic rejection of a rat trachea allograft was significantly inhibited by treatment with a selective PDGF inhibitor that blocks both and receptors (Kallio et al., 1999). An increased expression of PDGF A chain has also been noted in human cardiac allografts in myocytes and vascular structures with a predominance of the long form of the PDGF A chain (Zhao et al., 1995). In a rat model of heterotopic cardiac and aortic allografts, a PDGF receptor tyrosine kinase inhibitor, CGP53716, more selective for the PDGF receptor, reduced the incidence and severity of arteriosclerotic lesions (Sihvola et al., 1999). Other Fibrotic Responses As indicated in Table 2 and previously reviewed (Raines and Ross, 1993), PDGF and its receptors are increased in a number of inflammatory, fibroproliferative disorders, including inflammatory bowel disease, cirrhosis of the liver, myelofibrosis, progressive systemic sclerosis, and rheumatoid arthritis. As is suggested from recent studies of pulmonary fibrosis that implicate IL-1 (Bonner et al., 1998) and TNF in the regulation of PDGF and its receptors, it has recently been demonstrated that the endogenous production of IL-1 regulates PDGF A chain production in fibroblasts from patients with systemic sclerosis (Kawaguchi et al., 1999). PDGF in Tumors and in the Connective Tissue Stromal Response A large number of tumors and transformed cells have been shown to express PDGF and its receptors (see Table 3). Data for a causative role of PDGF and its receptors in tumorigenesis is most convincing in examples of altered expression as a result of the chromosomal rearrangement of either PDGF or its receptor. Deregulation of the PDGF B chain via fusion with the collagen gene COL1A1 has been demonstrated in dermatofibrosarcoma protuberans and giant cell fibroblastoma (Simon et al., 1997). Some of the most malignant tumors of neuroglial origin have been shown to contain amplification of the PDGF receptor (Smits and Funa, 1998). The TEL/PDGF receptor fusion product in chronic myelomonocytic leukemia is a transforming protein that self-associates and activates PDGF receptor kinase-dependent signaling pathways (Golub et al., 1994; Carroll et al., 1996). Although arguments have been put forward for an autocrine role of PDGF in the growth of a number of tumor cells (reviewed in Heldin and Westermark,
776 E. W. Raines 1999), it has not been determined whether it plays a critical role in tumor growth and progression. A more convincing argument can, however, be made for a role for PDGF in stimulation of the connective tissue stromal response (Raines and Ross, 1993; Heldin and Westermark, 1999). Stroma formation in nude mice inoculated with non-small cell lung cancer cells expressing elevated levels of PDGF, TGF , and TGF demonstrated prominent fibrous stroma, while those negative for PDGF, TGF , and TGF showed no significant stromal response (Bergh, 1988). Similarly, connective tissue is absent in mice following injection of WM9 melanoma cells devoid of PDGF B chain, while injection of melanoma cells transfected with the B chain cDNA showed abundant connective tissue septa surrounding the tumors (Forsberg et al., 1993).
IN THERAPY
Preclinical ± How does it affect disease models in animals? As detailed above, a number of animal models support a role for PDGF, particularly in connective tissue cell recruitment, in atherosclerosis and restenosis, glomerulonephritis, proliferative retinal disease, transplant rejection, pulmonary fibrosis, and other fibrotic responses, including the stromal response to tumors.
Effects of therapy: Cytokine, antibody to cytokine inhibitors, etc. The infusion of PDGF into animals and the inhibition of PDGF with blocking antibodies to PDGF and its receptors support the concept that PDGF stimulates SMC recruitment, proliferation, and extracellular matrix synthesis in a number of target organs (see the section on Pathophysiological roles).
Pharmacokinetics The intravenous injection of both unlabeled and [I125]PDGF-AB into baboons (Bowen-Pope et al., 1994a) and [I125]PDGF-BB into mice (Cohen et al., 1990) demonstrates rapid clearance with a half-life of less than 2 minutes. This rapid clearance is the result of widespread tissue distribution, metabolism by the liver, and excretion by the kidneys (Cohen et al., 1990). PDGF-AB implanted into subcutaneous wound chambers in rats (Sprugel et al., 1987) and
intraperitoneally, intramuscularly, or subcutaneously in mice (Abdiu et al., 1998) was also cleared rapidly, with a half-life of 12 hours. Maximal detection of PDGF in the blood was observed 2±4 hours after extravascular administration (Abdiu et al., 1998).
Toxicity No toxicity has been observed in clinical trials of PDGF for the treatment of lower extremity ulcers in diabetes (Smiell, 1998).
Clinical results PDGF is currently in use to accelerate the healing of chronic wounds (LeGrand, 1998; Robson et al., 1998; Miller, 1999; Rees et al., 1999). Although PDGF augments granulation tissue formation, fibroblast infiltration, and matrix formation in animal models of defective wound repair (see above), its efficacy in the treatment of human wounds has been limited and may require an approach utilizing a combination of growth factors (Robson et al., 1998).
References Abboud, H. E. (1995). Role of platelet-derived growth factor in renal injury. Annu. Rev. Physiol. 57, 297±309. Abdiu, A., Walz, T. M., and Wasteson, AÊ. (1998). Uptake of 125IPDGF-AB to the blood after extravascular administration in mice. Life Sci. 62, 1911±1918. Aird, W. C., Edelberg, J. M., Weiler-Guettler, H., Simmons, W. W., Smith, T. W., and Rosenberg, R. D. (1997). Vascular bedspecific expression of an endothelial cell gene is programmed by the tissue microenvironment. J. Cell Biol. 138, 1117±1124. Albelda, S. M., Elias, J. A., Levine, E. M., and Kern, J. A. (1989). Endotoxin stimulates platelet-derived growth factor production from cultured human pulmonary endothelial cells. Am. J. Physiol. 257, L65±L70. Alexander, R. J., Panja, A., Kaplan-Liss, E., Mayer, L., and Raicht, R. F. (1995). Expression of growth factor receptorencoded mRNA by colonic epithelial cells is altered in inflammatory bowel disease. Dig. Dis. Sci. 40, 485±494. Alitalo, R., Andersson, L. C., Betsholtz, C., Nilsson, K., Westermark, B., Heldin, C.-H., and Alitalo, K. (1987). Induction of platelet-derived growth factor gene expression during megakaryoblastic and monocytic differentiation of human leukemia cell lines. EMBO J. 6, 1213±1218. Alpers, C. E., Seifert, R. A., Hudkins, K. L., Johnson, R. J., and Bowen-Pope, D. F. (1992). Developmental patterns of PDGF B-chain, PDGF receptor, and alpha-actin expression in human glomerulogenesis. Kidney Int. 42, 390±399. Alpers, C. E., Hudkins, K. L., Ferguson, M., Johnson, R. J., and Rutledge, J. C. (1995). Platelet-derived growth factor A-chain expression in developing and mature human kidneys and in Wilms' tumor. Kidney Int. 48, 146±154. Alpers, C. E., Davis, C. L., Barr, D., Marsh, C. L., and Hudkins, K. L. (1996). Identification of platelet-derived growth
PDGF 777 factor A and B chains in human renal vascular rejection. Am. J. Pathol. 148, 439±451. Amur-Umarjee, S., Schonmann, V., and Campagnoni, A. T. (1997). Neuronal regulation of myelin basic protein mRNA translocation in oligodendrocytes is mediated by plateletderived growth factor. Dev. Neurosci. 19, 143±151. Andersson, M., OÈstman, A., Westermark, B., and Heldin, C.-H. (1994). Characterization of the retention motif in the C-terminal part of the long splice form of platelet-derived growth factor Achain. J. Biol. Chem. 269, 926±930. Andersson, M., OÈstman, A., Kreysing, J., BaÈckstroÈm, G., van de Poll, M., and Heldin, C.-H. (1995). Involvement of loop 2 of platelet-derived growth factor-AA and -BB in receptor binding. Growth Factors 12, 159±164. Andresen, J. L., and Ehlers, N. (1998). Chemotaxis of human keratocytes is increased by platelet-derived growth factor-BB, epidermal growth factor, transforming growth factoralpha, acidic fibroblast growth factor, insulin-like growth factor-I, and transforming growth factor-beta. Curr. Eye Res. 17, 79±87. Andrew, J. G., Hoyland, J. A., Freemont, A. J., and Marsh, D. R. (1995). Platelet-derived growth factor expression in normally healing human fractures. Bone 16, 455±460. Andrews, A., Balciunaite, E., Leong, F. L., Tallquist, M., Soriano, P., Refojo, M., and Kazlauskas, A. (1999). Plateletderived growth factor plays a key role in proliferative vitreoretinopathy. Invest. Ophthalmol. Vis. Sci. 40, 2683±2689. Ansel, J. C., Tiesman, J. P., Olerud, J. E., Krueger, J. G., Krane, J. F., Tara, D. C., Shipley, G. D., Gilbertson, D., Usui, M. L., and Hart, C. E. (1993). Human keratinocytes are a major source of cutaneous platelet-derived growth factor. J. Clin. Invest. 92, 671±678. Antoniades, H. N. (1981). Human platelet-derived growth factor (PDGF): purification of PDGF-I and PDGF-II and separation of their reduced subunits. Proc. Natl. Acad. Sci. USA 78, 7314± 7317. Antoniades, H. N., and Williams, L. T. (1983). Human plateletderived growth factor: structure and function. Fed. Proc. 42, 2630±2634. Ataliotis, P., and Mercola, M. (1997). Distribution and functions of platelet-derived growth factors and their receptors during embryogenesis. Int. Rev. Cytol. 172, 95±127. Balk, S. D. (1971). Calcium as a regulator of the proliferation of normal but not of transformed chicken fibroblasts in a plasma-containing medium. Proc. Natl. Acad. Sci. USA 68, 271±275. Bao, X., Lu, C., and Frangos, J. A. (1999). Temporal gradient in shear but not steady shear stress induces PDGF-A and MCP-1 expression in endothelial cells: role of NO, NF-B, and egr-1. Arterioscler. Thromb. Vasc. Biol. 19, 996±1003. Bar, R. S., Boes, M., Booth, B. A., Dake, B. L., Henley, S., and Hart, M. N. (1989). The effects of platelet-derived growth factor in cultured microvessel endothelial cells. Endocrinology 124, 1841±1848. Barnhill, R. L., Xiao, M., Graves, D., and Antoniades, H. N. (1996). Expression of the platelet-derived growth factor (PDGF)-A, PDGF-B and the PDGF-alpha receptor, but not the PDGF-beta receptor, in human malignant melanoma in vivo. Br. J. Dermatol. 135, 898±904. Barres, B. A., Hart, I. K., Coles, H. S., Burne, J. F., Voyvodic, J. T., Richardson, W. D., and Raff, M. C. (1992). Cell death and control of cell survival in the oligodendrocyte lineage. Cell 70, 31±46. Barrett, T. B., and Benditt, E. P. (1988). Platelet-derived growh factor gene expression in human atherosclerotic plaques and normal artery wall. Proc. Natl Acad. Sci. USA 85, 2810±2814.
Bassett, J. E., Bowen-Pope, D. F., Takayasu, M., and Dacey Jr, R. G. (1988). Platelet-derived growth factor does not constrict rat intracerebral arterioles in vitro. Microvasc. Res. 35, 368±373. Battegay, E. J., Raines, E. W., Seifert, R. A., Bowen-Pope, D. F., and Ross, R. (1990). TGF- induces bimodal proliferation of connective tissue cells via complex control of an autocrine PDGF loop. Cell 63, 515±524. Beck, P. L., and Podolsky, D. K. (1999). Growth factors in inflammatory bowel disease. Inflamm. Bowel Dis. 5, 44±60. Beer, H. D., Longaker, M. T., and Werner, S. (1997). Reduced expression of PDGF and PDGF receptors during impaired wound healing. J. Invest. Dermatol. 109, 132±138. Beitz, J. G., Kim, I. S., Calabresi, P., and Frackelton Jr, A. R. (1991). Human microvascular endothelial cells express receptors for platelet-derived growth factor. Proc. Natl Acad. Sci. USA 88, 2021±2025. Bergh, J. (1988). The expression of the platelet-derived and transforming growth factor genes in human nonsmall lung cancer cell lines is related to tumor stroma formation in nude mice tumors. Am. J. Pathol. 133, 434±439. Berk B. C., Alexander, R. W., Brock, T. A., Gimbrone Jr, M. A., and Webb, R. C. (1986). Vasoconstriction: a new activity for platelet-derived growth factor. Science 232, 87±90. Bernstein, J., Sella, O., Le, S.-Y., and Elroy-Stein, O. (1997). PDGF2/c-sis mRNA leader contains a differentiation-linked ribosomal entry site (D-IRES). J. Biol. Chem. 272, 9356±9362. Betsholtz, C. (1995). Role of platelet-derived growth factors in mouse development. Int. J. Dev. Biol. 39, 817±825. Betsholtz, C., and Raines, E. W. (1997). Platelet-derived growth factor: a key regulator of connective tissue cells in embryogenesis and pathogenesis. Kidney Int. 51, 1361±1369. Betsholtz, C., Johnsson, A., Heldin, C.-H., Westermark, B., Lind, P., Urdea, M. S., Eddy, R., Shows, T. B., Philpott, K., Mellor, A. L., Knott, T. J., and Scott, J. (1986). The human platelet-derived growth factor A-chain: complementary DNA sequence, chromosomal localization and expression in tumour cell lines. Nature 323, 226±232. Betsholtz, C., Bergh, J., Bywater, M., Pettersson, M., Johnsson, A., Heldin, C.-H., Ohlsson, R., Knott, T. J., Scott, J., Bell, G. I., and Westermark, B. (1987). Expression of multiple growth factors in a human lung cancer cell line. Int. J. Cancer 39, 502±507. Bidwell, M. C., Eitzman, B. A., Walmer, D. K., McLachlan, J. A., and Gray, K. D. (1995). Analysis of messenger ribonucleic acid and protein for the ligands and receptors of the platelet-derived growth factor signaling pathway in the placenta, extraembryonic membranes, and uterus during the latter half of murine gestation. Endocrinology 136, 5189±5201. Billett, M. A., Adbeish, I. S., Alrokayan, S. A., Bennett, A. J., Marenah, C. B., and White, D. A. (1996). Increased expression of genes for platelet-derived growth factor in circulating mononuclear cells of hypercholesterolemic patients. Arterioscler. Thromb. Vasc. Biol. 16, 399±406. Black, P. M., Carroll, R., Glowacka, D., Riley, K., and Dashner, K. (1994). Platelet-derived growth factor expression and stimulation in human meningiomas. J. Neurosurg. 81, 388± 393. Black, P., Carroll, R., and Glowacka, D. (1996). Expression of platelet-derived growth factor transcripts in medulloblastomas and ependymomas. Pediatr. Neurosurg. 24, 74±78. Boehm, K. D., Daimon, M., Gorodeski, I. G., Sheean, L. A., Utian, W. H., and Ilan, J. (1990). Expression of the insulinlike and platelet-derived growth factor genes in human uterine tissues. Mol. Reprod. Dev. 27, 93±101. Bonner, J. C., and Osornio-Vargas, A. R. (1995). Differential binding and regulation of platelet-derived growth factor A
778 E. W. Raines and B chain isoforms by alpha 2-macroglobulin. J. Biol. Chem. 270, 16236±16242. Bonner, J. C., Lindroos, P. M., Rice, A. B., Moomaw, C. R., and Morgan, D. L. (1998). Induction of PDGF receptor-alpha in rat myofibroblasts during pulmonary fibrogenesis in vivo. Am. J. Physiol. 274, L72±L80. Bonthron D. T., Morton, C. C., Orkin, S. H., and Collins, T. (1988). Platelet-derived growth factor A chain: gene structure, chromosomal location, and basis for alternative mRNA splicing. Proc. Natl Acad. Sci. USA 85, 1492±1496. BostroÈm, H., Willetts, K., Pekny, M., LeveÂen, P., Lindahl, P., Hedstrand, H., Pekna, M., HellstroÈm, M., Gebre-Medhin, S., Schalling, M., Nilsson, M., Kurland, S., Tornell, J., Heath, J. K., and Betsholtz, C. (1996). PDGF-A signaling is a critical event in lung alveolar myofibroblast development and alveogenesis. Cell 85, 863±873. Bowen-Pope, D. F., and Ross, R. (1985). In ``Methods in Enzymology, Hormone Action, Part I: Peptide Hormones, Vol. 109'' (ed L. Birnbaumer and B. W. O'Malley), Methods for studying the platelet-derived growth factor receptor, pp. 69± 100. Academic Press, New York. Bowen-Pope, D. F., Malpass, T. W., Foster, D. M., and Ross, R. (1984a). Platelet-derived growth factor in vivo: levels, activity, and rate of clearance. Blood 64, 458±469. Bowen-Pope, D. F., Vogel, A., and Ross, R. (1984b). Production of platelet-derived growth factor-like molecules and reduced expression of platelet-derived growth factor receptors accompany transformation by a wide spectrum of agents. Proc. Natl Acad. Sci. USA 81, 2396±2400. Bowen-Pope, D. F., Hart, C. E., and Seifert, R. A. (1989). Sera and conditioned media contain different isoforms of plateletderived growth factor (PDGF) which bind to different classes of PDGF receptor. J. Biol. Chem. 264, 2502±2508. Breslow, J. L. (1997). Cardiovascular disease burden increases, NIH funding decreases. Nature Med. 3, 600±601. Brewitt, B., and Clark, J. I. (1988). Growth and transparency in the lens, an epithelial tissue, stimulated by pulses of PDGF. Science 242, 777±779. Bronzert, D. A., Pantazis, P., Antoniades, H. N., Kasid, A., Davidson, N., Dickson, R. B., and Lippman, M. E. (1987). Synthesis and secretion of platelet-derived growth factor by human breast cancer cell lines. Proc. Natl Acad. Sci. USA 84, 5763±5767. Bronzert, D. A., Bates, S. E., Sheridan, J. P., Lindsey, R., Valverius, E. M., Stampfer, M. R., Lippman, M. E., and Dickson, R. B. (1990). Transforming growth factor-beta induces platelet-derived growth factor (PDGF) messenger RNA and PDGF secretion while inhibiting growth in normal human mammary epithelial cells. Mol. Endocrinol. 4, 981± 989. Brown, D. M., Hong, S. P., Farrell, C. L., Pierce, G. F., and Khouri, R. K. (1995). Platelet-derived growth factor BB induces functional vascular anastomoses in vivo. Proc. Natl Acad. Sci. USA 92, 5920±5924. Campochiaro, P. A., and Glaser, B. M. (1985). Platelet-derived growth factor is chemotactic for human retinal pigment epithelial cells. Arch. Ophthalmol. 103, 576±579. Campochiaro, P. A., Sugg, R., Grotendorst, G., and Hjelmeland, L. M. (1989). Retinal pigment epithelial cells produce PDGF-like proteins and secrete them into their media. Exp. Eye Res. 49, 217±227. Carroll, M., Tomasson, M. H., Barker, G. F., Golub, T. R., and Gilliland, D. G. (1996). The TEL/platelet-derived growth factor beta recptor (PDGF beta R) fusion in chronic myelomonocytic leukemia is a transforming protein that self-associates and
activates PDGF beta R kinase-dependent signaling pathways. Proc. Natl Acad. Sci. USA 93, 14845±14850. Centrella, M., McCarthy, T. L., Kusmik, W. F., and Canalis, E. (1991). Relative binding and biochemical effects of heterodimeric and homodimeric isoforms of platelet-derived growth factor in osteoblast-enriched culture from fetal rat bone. J. Cell. Physiol. 147, 420±426. Chai, Y., Bringas, P. Jr., Mogharei, A., Shuler, C. F., and Slavkin, H. C. (1998). PDGF-A and PDGFR-alpha regulate tooth formation via autocrine mechanism during mandibular morphogenesis in vitro. Dev. Dyn. 213, 500±511. Chung, C. K., and Antoniades, H. N. (1992). Expression of c-sis/ platelet-derived growth factor B, insulin-like growth factor I, and transforming growth factor alpha messenger RNAs and their respective receptor messenger RNAs in primary human gastric carcinomas: in vivo studies with in situ hybridization and immunocytochemistry. Cancer Res. 52, 3453±3459. Cochran, B. H., Zullo, J., Verma, I. M., and Stiles, C. D. (1984). Expression of the c-fos gene and of a fos-related gene is stimulated by platelet-derived growth factor. Science 226, 1080±1082. Cohen, A. M., Soderberg, C., and Thomason, A. (1990). Plasma clearance and tissue distribution of recombinant human platelet-derived growth factor (B-chain homodimer) in rats. J. Surg. Res. 49, 447±452. Collins, T., Bonthron, D. T., and Orkin, S. H. (1987a). Alternative RNA splicing affects function of encoded platelet-derived growth factor A chain. Nature 328, 621±624. Collins, T., Pober, J. S., Gimbrone Jr, M. A., Hammacher, A., Betsholtz, C., Westermark, B., and Heldin, C.-H. (1987b). Cultured human endothelial cells express platelet-derived growth factor A chain. Am. J. Pathol. 126, 7±12. Coltrera, M. D., Wang, J., Porter, P. L., and Gown, A. M. (1995). Expression of platelet-derived growth factor B-chain and the platelet-derived growth factor receptor beta subunit in human breast tissue and breast carcinoma. Cancer Res. 55, 2703±2708. Conn, G., Bayne, M. L., Soderman, D. D., Kwok, P. W., Sullivan, K. A., Palisi, T. M., Hope, D. A., and Thomas, K. A. (1990). Amino acid and cDNA sequences of a vascular endothelial cell mitogen that is homologous to plateletderived growth factor. Proc. Natl Acad. Sci. USA 87, 2628± 2632. Crosby, J. R., Seifert, R. A., Soriano, P., and Bowen-Pope, D. F. (1998). Chimaeric analysis reveals role of PDGF receptors in all muscle lineages. Nature Genet. 18, 385±388. Crosby, J. R., Tappan, K. A., Seifert, R. A., and BowenPope, D. F. (1999). Chimera analysis reveals that fibroblasts and endothelial cells require platelet-derived growth factor receptor beta expression for participation in reactive connective tissue formation in adults but not during development. Am. J. Pathol. 154, 1315±1321. Cunningham, L. D., Brecher, P., and Cohen, R. A. (1992). Platelet-derived growth factor receptors on macrovascular endothelial cells mediate relaxation via nitric oxide in the rat aorta. J. Clin. Invest. 89, 878±882. Dainiak, N., Davies, G., Kalmanti, M.., Lawler, J., and Kulkarni, V. (1983). Platelet-derived growth factor promotes proliferation of erythropoietic progenitor cells in vitro. J. Clin. Invest. 71, 1206±1214. Dalla Favera, R., Gallo, R. C., Giallongo, A., and Croce, C. M. (1982). Chromosomal localization of the human homolog (c-sis of the simian sarcoma virus onc gene). Science 218, 686± 688. Daniel, T. O., Gibbs, V. C., Milfay, D. F., Garovoy, M. R., and Williams, L. T. (1986). thrombin stimulates c-sis gene
PDGF 779 expression in microvascular endothelial cells. J. Biol. Chem. 261, 9579±9582. Davis, J. B., and Stroobant, P. (1990). Platelet-derived growth factors and fibroblast growth factors are mitogens for rat Schwann cells. J. Cell Biol. 110, 1353±1360. Daynes, R. A., Dowell, T., and Araneo, B. A. (1991). Plateletderived growth factor is a potent biologic response modifier of T cells. J. Exp. Med. 174, 1323±1333. Delwiche, F., Raines, E., Powell, J., Ross, R., and Adamson, J. (1985). Platelet-derived growth factor enhances in vitro erythropoiesis via stimulation of mesenchymal cells. J. Clin. Invest. 76, 137±142. Deuel, T. F., Huang, J. S., Proffitt, R. T., Baenziger, J. U., Chang, D., and Kennedy, B. B. (1981). Human platelet-derived growth factor ± purification and resolution into two active protein fractions. J. Biol. Chem. 256, 8896±8899. Deuel, T. F., Senior, R. M., Huang, J. S., and Griffin, G. L. (1982). Chemotaxis of monocytes and neutrophils to plateletderived growth factor. J. Clin. Invest. 69, 1046±1049. DiCorleto, P. E., and Bowen-Pope, D. F. (1983). Cultured endothelial cells produce a platelet-derived growth factor-like protein. Proc. Natl Acad. Sci. USA 80, 1919±1923. Dirks, R. P., Jansen, H. J., Gerritsma, J., Onnekink, C., and Bloemers, H. P. (1993). Localization and functional analysis of DNase-1-hypersensitive sites in the human c-sis/PDGF-B gene transcription unit and its flanking regions. Eur. J. Biochem. 211, 509±519. Dirks, R. P. H., Jansen, H. J., van Gerven, B., Onnekink, C., and Bloemers, H. P. J. (1995). In vivo footprinting and functional analysis of the human c-sis/PDGF B gene promoter provides evidence for two binding sites for transcriptional activators. Nucleic Acids Res. 23, 1119±1126. Di Rocco, F., Carroll, R. S., Zhang, J., and Black, P. M. (1998). Platelet-derived growth factor and its receptor expression in human oligodendrogliomas. Neurosurg. 42, 341±346. Dobrian, A., Wade, S. S., and Prewitt, R. L. (1999). PDGF-A expression correlates with blood pressure and remodeling in 1K1C hypertensive rat arteries. Am. J. Physiol. 276, H2159± H2167. Durante, W., Liao, L., Iftikhar, I., Cheng, K., and Schafer, A. I. (1996). Platelet-derived growth factor regulates vascular smooth muscle cell proliferation by inducing cationic amino acid transporter gene expression. J. Biol. Chem. 271, 11838±11843. Dvonch, V. M., Murphey, R. J., Matsuoka, J., and Grotendorst, G. R. (1992). Changes in growth factor levels in human wound fluid. Surgery 112, 18±23. Ebert, M., Yokoyama, M., Friess, H., Kobrin, M. S., Buchler, M. W., and Korc, M. (1995). Induction of plateletderived growth factor A and B chains and over-expression of the receptors in human pancreatic cancer. Int. J. Cancer 62, 529±535. Ebert, M., Kasper, H. U., Hernberg, S., Friess, H., Buchler, M. W., Roessner, A., Korc, M., and Malfertheiner, P. (1998). Overexpression of platelet-derived growth factor (PDGF) B chain and type beta PDGF receptor in human chronic pancreatitis. Dig. Dis. Sci. 43, 567±574. Edelberg, J. M., Aird, W. C., Wu, W., Rayburn, H., Mamuya, W. S., Mercola, M., and Rosenberg, R. D. (1998). PDGF mediates cardiac microvascular communication. J. Clin. Invest. 102, 837±843. Eming, S. A., Lee, J., Snow, R. G., Tompkins, R. G., Yarmush, M. L., and Morgan, J. R. (1995). Genetically modified human epidermis overexpressing PDGF-A directs the development of a cellular and vascular connective tissue stroma when transplanted to athymic mice ± implications for the use of
genetically modified keratinocytes to modulate dermal regeneration. J. Invest. Dermatol. 105, 756±763. Eming, S. A., Medalie, D. A., Tompkins, R. G., Yarmush, M. L., and Morgan, J. R. (1998). Genetically modified human keratinocytes overexpressing PDGF-A enhance the performance of a composite skin graft. Hum. Gene Ther. 9, 529±539. Eming, S. A., Yarmush, M. L., Krueger, G. G., and Morgan, J. R. (1999). Regulation of the spatial organization of mesenchymal connective tissue: effects of cell-associated versus released isoforms of platelet-derived growth factor. Am. J. Pathol. 154, 281±289. Ernofsson, M., and Siegbahn, A. (1996). Platelet-derived growth factor-BB and monocyte chemotactic protein-1 induce human peripheral blood monocytes to express tissue factor. Thromb. Res. 83, 307±320. Evanko, S. P., Raines, E. W., Ross, R., Gold, L. I., and Wight, T. N. (1998). Proteoglycan distribution in lesions of atherosclerosis depends on lesion severity, structural characteristics, and the proximity of platelet-derived growth factor and transforming growth factor- . Am. J. Pathol. 152, 533±546. Fahrer, C., Brachmann, R., and von der Helm, K. (1989). Expression of c-sis and other cellular proto-oncogenes in human sarcoma cell lines and biopsies. Int. J. Cancer 44, 652± 657. FellstroÈm, B., Klareskog, L., Heldin, C.-H., Larsson, E., RoÈnnstrand, L., Terracio, L., Tufveson, G., Wahlberg, J., and Rubin, K. (1989). Platelet-derived growth factor receptors in the kidney ± upregulated expression in inflammation. Kidney Int. 36, 1099±1102. Fen, Z., and Daniel, T. O. (1991). 50 untranslated sequences determine degradative pathway for alternate PDGF B/c-sis mRNA's. Oncogene 6, 953±959. Ferns, G. A. A., Sprugel, K. H., Seifert, R. A., Bowen-Pope, D. F., Kelly, J. D., Murray, M., Raines, E. W., and Ross, R. (1990). Relative platelet-derived growth factor receptor subunit expression determines cell migration to different dimeric forms of PDGF. Growth Factors 3, 315±324. Feyzi, E., Lustig, F., Fager, G., Spillmann, D., Lindahl, U., and Salmivirta, M. (1997). Characterization of heparin and heparan sulfate domains binding to the long splice variant of platelet-derived growth factor A chain. J. Biol. Chem. 272, 5518±5524. Fischer, W. H., and Schubert, D. (1996). Characterization of a novel platelet-derived growth factor-associated protein. J. Neurochem. 66, 2213±2216. Floege, J., Eng, E., Young, B. A., Alpers, C. E., Barrett, T. B., Bowen-Pope, D. F., and Johnson, R. J. (1993). Infusion of platelet-derived growth factor or basic fibroblast growth factor induces selective glomerular mesangial cell proliferation and matrix accumulation in rats. J. Clin. Invest. 92, 2952±2962. Floege, J., Hudkins, K. L., Davis, C. L., Schwartz, S. M., and Alpers, C. E. (1998). Expression of PDGF alpha-receptor in renal arteriosclerosis and rejecting renal transplants. J. Am. Soc. Nephrol. 9, 211±223. Floege, J., Ostendorf, T., Janssen, J., Burg, M., Radeke, H. H., Vargeese, C., Gill, S. D., Green, L. S., and Janjic, N. (1999). Novel approach to specific growth factor inhibition in vivo: antagonism of platelet-derived growth factor in glomerulonephritis by aptamers. Am. J. Pathol. 154, 169±179. Forsberg, K., Valyi-Nagy, I., Heldin, C.-H., Herlyn, M., and Westermark, B. (1993). Platelet-derived growth factor 7(PDGF) in oncogenesis: development of a vascular connective tissue stroma in xenotransplanted human melanoma producing PDGF-BB. Proc. Natl Acad. Sci. USA 90, 393±397.
780 E. W. Raines Forsberg-Nilsson, K., Behar, T. N., Afrakhte, M., Barker, J. L., and McKay, R. D. (1998). Platelet-derived growth factor induces chemotaxis of neuroepithelial stem cells. J. Neurosci. Res. 53, 521±530. Fraizer, G. E., Bowen-Pope, D. F., and Vogel, A. M. (1987). Production of platelet-derived growth factor by cultured Wilms' tumor cells and fetal kidney cells. J. Cell. Physiol. 133, 169±174. Franchini, A., Kletsas, D., and Ottaviani, E. (1996). Immunocytochemical evidence of PDGF- and TGF-beta-like molecules in invertebrate and vertebrate immunocytes: an evolutionary approach. Histochem. J. 28, 599±605. Francis, S. E., Hunter, S., Holt, C. M., Gadsdon, P. A., Rogers, S., Duff, G. W., Newby, A. C., and Angelini, G. D. (1994). Release of platelet-derived growth factor from pig venous arterial grafts. J. Thorac. Cardiovasc. Surg. 108, 540±548. Franklin, G. C., Donovan, M., Adam G. I., Holmgren, L., PfeiferOhlsson, S., and Ohlsson, R. (1991). Expression of the human PDGF-B gene is regulated by both positively and negatively acting cell type-specific regulatory elements located in the first intron. EMBO J. 10, 1365±1373. Fruttiger, M., Calver, A. R., Kruger, W. H., Mudhar, H. S., Michalovich, D., Takakura, N., Nishikawa, S., and Richardson, W. D. (1996). PDGF mediates a neuron-astrocyte interaction in the developing retina. Neuron. 17, 1117±1131. Fruttiger, M., Karlsson, L., Hall, A. C., Abramsson, A., Calver, A. R., BostroÈm, H., Willetts, K., Bertold, C. H., Heath, J. K., Betsholtz, C., and Richardson, W. D. (1999). Defective oligodendrocyte development and severe hypomyelination in PDGF-A knockout mice. Development 126, 457±467. Fudge, K., Wang, C. Y., and Stearns, M. E. (1994). Immunohistochemistry analysis of platelet-derived growth factor A and B chains and platelet-derived growth factor alpha and beta receptor expression in benign prostatic hyperplasias and Gleason-graded human prostate adenocarcinomas. Mod. Pathol. 7, 549±554. Fujii, H., Kitazawa, R., Maeda, S., Mizuno, K., and Kitazawa, S. (1999). Expression of platelet-derived growth factor proteins and their receptor alpha and beta mRNAs during fracture healing in the normal mouse. Histochem. Cell Biol. 112, 131±138. Fukuda, N., Kubo, A., Watanabe, Y., Nakayama, T., Soma, M., Izumi, Y., and Kanmatsuse, K. (1997a). Antisense oligodeoxynucleotide complementary to platelet-derived growth factor A-chain messenger RNA inhibits the arterial proliferation in spontaneously hypertensive rats without altering their blood pressures. J. Hypertens. 15, 1123±1136. Fukuda, N., Kishioka, H., Satoh, C., Nakayama, T., Watanabe, Y., Soma, M., Izumi, Y., and Kanmatsuse, K. (1997b). Role of long-form PDGF A-chain in the growth of vascular smooth muscle cells from spontaneously hypertensive rats. Am. J. Hypertens. 10, 1117±1124. Gay, C. G., and Winkles, J. A. (1990). Heparin-binding growth factor-1 stimulation of human endothelial cells induces plateletderived growth factor A-chain gene expression. J. Biol. Chem. 265, 3284±3292. Gay, S., Jones, R.E. Jr., Huang, G. Q., and Gay, R. E. (1989). Immunohistologic demonstration of platelet-derived growth factor (PDGF) and sis-oncogene expression in scleroderma. J. Invest. Dermaol. 92, 301±303. Gersuk, G. M., Westermark, B., Mohabeer, A. J., Challita, P. M., Pattamakom, S., and Pattengale, P. K. (1991). Inhibition of human natural killer cell activity by platelet-derived growth factor (PDGF).III: Membrane binding studies and differential biological effects of recombinant PDGF isoforms. Scand. J. Immunol. 33, 521±532.
Gerwin, B. I., Lechner, J. F., Reddel, R. R., Roberts, A. B., Robbins, K. C., Gabrielson, E. W., and Harris, C. C. (1987). Comparison of production of transforming growth factor-beta and platelet-derived growth factor by normal human mesothelial cells and mesothelioma cell lines. Cancer Res. 47, 6180± 6184. Gesualdo, L., Di Paolo, S., Milani, S., Pinzani, M., Grappone, C., Ranieri, E., Pannarale, G., and Schena, F. P. (1994). Expression of platelet-derived growth factor receptors in normal and diseased human kidney. An immunohistochemistry and in situ hybridization study. J. Clin. Invest. 94, 50±58. Ghil, J. S., and Chung, H. M. (1999). Evidence that platelet derived growth factor (PDGF) action is required for mesoderm patterning in early amphibian (Xenopus laevis) embryogenesis. Int. J. Dev. Biol. 43, 329±334. Giacobini, M. M., Almstrom, S., Funa, K., and Olson, L. (1993). Differential effects of platelet-derived growth factor isoforms on dopamine neurons in vivo: -BB supports cell survival, -AA enhances fiber formation. Neurosci. 57, 923±929. Giese, N. A., Robbins, K. C., and Aaronson, S. A. (1987). The role of individual cysteine residues in the structure and function of the v-sis gene product. Science 236, 1315±1318. Giese, N. A., Marijianowski, M. M., McCook, O., Hancock, A., Ramakrishnan, V., Fretto, L. J., Chen, C., Kelly, A. B., Koziol, J. A., Wilcox, J. N., and Hanson, S. R. (1999). The role of alpha and beta platelet-derived growth factor receptor in the vascular response to injury in nonhuman primates. Arterioscler. Thromb. Vasc. Biol. 19, 900±909. Gillaspy, G. E., Mapstone, T. B., Samols, D., and Goldthwait, D. A. (1992). Transcriptional patterns of growth factors and proto-oncogenes in human glioblastomas and normal glial cells. Cancer Lett. 65, 55±60. Gnessi, L., Emidi, A., Farini, D., Scarpa, S., Modesti, A., Ciampani, T., Silvestroni, L., and Spera, G. (1992). Rat Leydig cells bind platelet-derived growth factor through specific receptors and produce platelet-derived growth factor-like molecules. Endocrinology 130, 2219±2224. Gnessi, L., Emidi, A., Scarpa, S., Palleschi, S., RaganoCaracciolo, M., Silvestroni, L., Modesti, A., and Spera, G. (1993). Platelet-derived growth factor effects on purified testicular peritubular myoid cells: binding, cytosolic Ca2 increase, mitogenic activity, and extracellular matrix production enhancement. Endocrinology 133, 1880±1890. Gnessi, L., Emidi, A., Jannini, E. A., Carosa, E., Maroder, M., Arizzi, M., Ulisse, S., and Spera, G. (1995). Testicular development involves the spatiotemporal control of PDGF's and PDGF receptors gene expression and action. J. Cell Biol. 131, 1105±1121. Gohring, W., Sasaki, T., Heldin, C.-H., and Timpl, R. (1998). Mapping of the binding of platelet-derived growth factor to distinct domains of the basement membrane proteins BM-40 and perlecan and distinction from the BM-40 collagen-binding epitope. Eur. J. Biochem. 255, 60±66. Golub, T. R., Barker, G. F., Lovett, M., and Gilliland, D. G. (1994). Fusion of PDGF receptor beta to a novel ets-like gene, tel, in chronic myelomonocytic leukemia with t(5;12) chromosomal translocation. Cell 77, 307±316. Goretzki, L., Burg, M. A., Grako, K. A., and Stallcup, W. B. (1999). High-affinity binding of basic fibroblast growth factor and platelet-derived growth factor-AA to the core protein of the NG2 proteoglycan. J. Biol. Chem. 274, 16831±16837. Goustin, A. S., Betsholtz, C., Pfeifer-Ohlsson, S., Persson, H., Rydnert, J., Bywater, M., Holmgren, G., Heldin, C.-H., Westermark, B., and Ohlsson, R. (1985). Coexpression of the sis and myc proto-oncogenes in developing human placenta
PDGF 781 suggests autocrine control of trophoblast growth. Cell 41, 301± 312. Goustin, A. S., Galanopoulos, T., Kalyanaraman, V. S., and Pantazis, P. (1990). Coexpression of the genes for plateletderived growth factor and its receptor in human T-cell lines infected with HTLV-I. Growth Factors 2, 189±195. Graves, D. T., Grotendorst, G. R., Antoniades, H. N., Schwartz, C. J., and Valente, A. J. (1989). Platelet-derived growth factor is not chemotactic for human peripherial blood monocytes. Exp. Cell Res. 180, 497±503. Gray, K., Eitzman, B., Raszmann, K., Steed, T., Geboff, A., McLachlan, J., and Bidwell, M. (1995). Coordinate regulation by diethylstilbestrol of the platelet-derived growth factor-A (PDGF-A) and -B chains and the PDGF receptor alpha- and beta-subunits in the mouse uterus and vagina: potential mediators of estrogen action. Endocrinology 136, 2325±2340. Green, R. S., Lieb, M. E., Weintraub, A. S., Gacheru, S. N., Rosenfield, C. L., Shah, S., Kagan, H. M., and Taubman, M. B. (1995). Identification of lysyl oxidase and other platelet-derived growth factor-inducible genes in vascular smooth muscle cells by differential screening. Lab. Invest. 73, 476±482. Greenberg, M. E., and Ziff, E. B. (1984). Stimulation of 3T3 cells induces transcription of the c-fos proto-oncogene. Nature 311, 433±438. Greenhalgh, D. G., Sprugel, K. H., Murray, M. J., and Ross, R. (1990). PDGF and FGF stimulate wound healing in the genetically diabetic mouse. Am. J. Pathol. 136, 1235±1246. Grotendorst, G. R., Martin, G. R., Pencev, D., Sodek, J., and Harvey, A. K. (1985). Stimulation of granulation tissue formation by platelet-derived growth factor in normal and diabetic rats. J. Clin. Invest. 76, 2323±2329. Guha, A., Glowacka, D., Carroll, R., Dashner, K., Black, P. M., and Stiles, C. D. (1995). Expression of platelet-derived growth factor and platelet-derived growth factor receptor mRNA in a glioblastoma from a patient with Li-Fraumeni syndrome. J. Neurol. Neurosurg. Psychiatry 58, 711±714. Guillausseau, P. J., Dupuy, E., Bryckaert, M. C., Timsit, J., Chanson, P. Tobelem, G., Caen, J. P., and Lubetzki, J. (1989). Platelet-derived growth factor (PDGF) in type 1 diabetes mellitus. Eur. J. Clin. Invest. 19, 172±175. Guillot, P. V., Guan, J., Liu, L., Kuivenhoven, J. A., Rosenberg, R. D., Sessa, W. C., and Aird, W. C. (1999). A vascular bed-specific pathway. J. Clin. Invest. 103, 799±805. Gurski, M. R., Gonzalez, E., and Brown, E. G. (1999). Immunochemical localization of platelet-derived growth factor in placenta and its possible role in pre-eclampsia. J. Investig. Med. 47, 128±133. Hall, A., Giese, N. A., and Richardson, W. D. (1996). Spinal cord oligodendrocytes develop from ventrally derived progenitor cells that express PDGF alpha-receptors. Development 122, 4085±4094. Hammacher, A., Hellman, U., Johnsson, A., OÈstman, A., Gunnarsson, K., Westermark, B., Wasteson, AÊ., and Heldin, C.-H. (1988a). A major part of the platelet-derived growth factor purified from human platelets is a heterodimer of one A and one B chain. J. Biol. Chem. 263, 16493±16498. Hammacher, A., NisteÂr, M., Westermark, B., and Heldin, C.-H. (1988b). A human glioma cell line secretes three structurally and functionally different dimeric forms of platelet-derived growth factor. Eur. J. Biochem. 176, 179±186. Hanai, K., Kato, H., Matsuhashi, S., Morita, H., Raines, E. W., and Ross, R. (1987). Platelet proteins, including platelet-derived growth factor, specifically depress a subset of the multiple components of the response elicited by glutathione in Hydra. J. Cell Biol. 104, 1675±1681.
Handa, J. T., Reiser, K. M., Matsunaga, H., and Hjelmeland, L. M. (1998). The advanced glycation endproduct pentosidine induces the expression of PDGF-B in human retinal pigment epithelial cells. Exp. Eye Res. 66, 411±419. Haniu, M., Rohde, M. F., and Kenney, W. C. (1993). Disulfide bonds in recombinant human platelet-derived growth factor BB dimer: characterization of intermolecular and intramolecular disulfide linkages. Biochemistry 32, 2431±2437. Hannink, M., and Donoghue, D. J. (1986). Cell surface expression of membrane-anchored v-sis gene products: glycosylation is not required for cell surface transport. J. Cell Biol. 103, 2311±2322. Hannink, M., and Donoghue, D. J. (1988). Autocrine stimulation by the v-sis gene product requires a ligand-receptor interaction at the cell surface. J. Cell Biol. 107, 287±298. Hardy, M., Reddy, U. R., and Pleasure, D. (1992). Plateletderived growth factor and regulation of Schwann cell proliferation in vivo. J. Neurosci. Res. 31, 254±262. Hart, C. E., Forstrom, J. W., Kelly, J. D., Seifert, R. A., Smith, R. A., Ross, R., Murray, M. J., and Bowen-Pope, D. F. (1988). Two classes of PDGF receptor recognize different isoforms of PDGF. Science 240, 1529±1531. Hart, C. E., Bailey, M., Curtis, D. A., Osborn, S., Raines, E., Ross, R., and Forstrom, J. W. (1990). Purification of PDGFAB and PDGF-BB from human platelet extracts and the identification of all three PDGF dimers in human platelets. Biochemistry 29, 166±172. Hauschka, P. V., Chen, T. L., and Macrakos, A. E. (1988). Polypeptide growth factors in bone matrix. Ciba Found. Symp. 136, 207±225. Heldin, C.-H., and Westermark, B. (1999). Mechanism of action and in vivo role of platelet-derived growth factor. Physiol. Rev. 79, 1283±1316. Heldin, C.-H., Westermark, B., and Wasteson, AÊ. (1979). Plateletderived growth factor: purification and partial characterization. Proc. Natl Acad. Sci. USA 76, 3722±3726. Heldin, C.-H., Westermark, B., and Wasteson, AÊ. (1981a). Platelet-derived growth factor: isolation by a large-scale procedure and analysis of subunit composition. Biochem. J. 193, 907± 913. Heldin, C.-H., Westermark, B., and Wasteson, AÊ. (1981b). Specific receptors for platelet-derived growth factor on cells derived from connective tissue and glia. Proc. Natl Acad. Sci. USA 78, 3664±3668. Heldin, C.-H., BaÈckstroÈm, OÈstman, A., Hammacher, A., Ronnstrand, L., Rubin, K., NisteÂr, M., and Westermark, B. (1988). Binding of different dimeric forms of PDGF to human fibroblasts: evidence for two separate receptor types. EMBO J. 7, 1387±1393. Heldin, P., Laurent, T. C., and Heldin, C.-H. (1989). Effect of growth factors on hyaluronan synthesis in cultured human fibroblasts. Biochem. J. 258, 919±922. Heldin, P., Pertoft, H., Nordlinder, H., Heldin, C.-H., and Laurent, T. C. (1991). Differential expression of platelet-derived growth factor alpha- and beta-receptors on fat storing cells and endothelial cells of rat liver. Exp. Cell Res. 193, 364±369. HellstroÈm, M., Kalen, M., Lindahl, P., Abramsson, A., and Betsholtz, C. (1999). Role of PDGF-B and PDGFR-beta in recruitment of vascular smooth muscle cells and pericytes during embryonic blood vessel formation in the mouse. Development 126, 3047±3055. Herbert, J. M., Lamarche, I., and Carmeliet, P. (1997). Urokinase and tissue-type plasminogen activator are required for the mitogenic and chemotactic effects of bovine fibroblast growth factor and platelet-derived growth factor-BB for vascular smooth muscle cells. J. Biol. Chem. 272, 23585±23591.
782 E. W. Raines Heuchel, R., Berg, A., Tallquist, M., Ahlen, K., Reed, R. K., Rubin, K., Claesson-Welsh, L., Heldin, C.-H., and Soriano, P. (1999). Platelet-derived growth factor beta receptor regulates interstitial fluid homeostasis through phosphatidylinositol-30 kinase signaling. Proc. Natl Acad. Sci. USA 96, 11410±11415. Hock, J. M., and Canalis, E. (1994). Platelet-derived growth factor enhances bone cell replication, but not differentiated function of osteoblasts. Endocrinology 134, 1423±1428. Hogg, P. J., Hotchkiss, K. A., Jimenez, B. M., Stathakis, P., and Chesterman, C. N. (1997). Interaction of platelet-derived growth factor with thrombospondin 1. Biochem. J. 326, 709± 716. Horvath, P., Suganuma, A., Inaba, M., Pan, Y. B., and Gupta, K. C. (1995). Multiple elements in the 50 untranslated region down-regulate c-sis messenger RNA translation. Cell Growth Differ. 6, 1103±1110. Hossain, M. Z., Jagdale, A. B., Ao, P., Kazlauskas, A., and Boynton, A. L. (1999). Disruption of gap junctional communication by the platelet-derived growth factor is mediated via multiple signaling pathways. J. Biol. Chem. 274, 10489±10496. Hoyle, G. W., Li, J., Finkelstein, J. B., Eisenberg, T., Liu, J. Y., Lasky, J. A., Athas, G., Morris, G. F., and Brody, A. R. (1999). Emphysematous lesions, inflammation, and fibrosis in the lungs of transgenic mice overexpressing platelet-derived growth factor. Am. J. Pathol. 154, 1763±1775. Huang, J. S., Huang, S. S., and Deuel, T. F. (1984). Specific covalent binding of platelet-derived growth factor to human plasma alpha 2-macroglobulin. Proc. Natl Acad. Sci. USA 81, 342±346. Iida, H., Seifert, R., Alpers, C. E., Gronwald, R. G., Phillips, P. E., Pritzl, P., Gordon, K., Gown, A. M., Ross, R., BowenPope, D. F., and Johnson, R. J. (1991). Platelet-derived growth factor (PDGF) and PDGF receptor are induced in mesangial proliferative nephritis in the rat. Proc. Natl Acad. Sci. USA 88, 6560±6564. Iihara, K., Hashimoto, N., Tsukahara, T., Sakata, M., Yanamoto, H., and Taniguchi, T. (1997). Platelet-derived growth factor-BB, but not -AA, prevents delayed neuronal death after forebrain ischemia in rats. J. Cereb. Blood Flow Metab. 17, 1097±1106. Ikeda, M., Morita, C., Mizuno, M., Sada, T., Koike, H., and Kurokawa, K. (1997). PDGF-BB decreases systolic blood pressure through an increase in macrovascular compliance in rats. Am. J. Physiol. 273, H1719±H1726. Iwasaki, Y., Shiojima, T., Tagaya, N., Kobayashi, T., and Kinoshita, M. (1997). Basic fibroblast growth factor and platelet-derived growth factor prevent the death of spinal motor neurons after sciatic nerve transection in the neonatal rats. Neurol. Res. 19, 555±557. Jin, P., Rahm, M., Claesson-Welsh, L., Heldin, C.-H., and Sejersen, T. (1990). Expression of PDGF A-chain and betareceptor genes during rat myoblast differentiation. J. Cell Biol. 110, 1665±1672. Jin, P., Sejersen, T., and Ringertz, N. R. (1991). Recombinant platelet-derived growth factor-BB stimulates growth and inhibits differentiation of rat L6 myoblasts. J. Biol. Chem. 266, 1245±1249. Johnson, R. J., Raines, E. W., Floege, J., Yoshimura, A., Pritzl, P., Alpers, C., and Ross, R. (1992). Inhibition of mesangial cell proliferation and matrix expansion in gomerulonephritis in the rat by antibody to platelet-derived growth factor. J. Exp. Med. 175, 1413±1416. Johnsson, A., Heldin, C.-H., Wasteson, AÊ., Westermark, B., Deuel, T. F., Huang, J. S., Seeburg, P. H., Gray, A., Ullrich, A., Scrace, G., Stroobant, P., and Waterfield, M. D. (1984). The
c-sis gene encodes a precursor of the B chain of platelet-derived growth factor. EMBO J. 3, 921±928. Jones, S. D., Ho, L., Smith, J. C., Yordan, C., Stiles, C. D., and Mercola, M. (1993). The Xenopus platelet-derived growth factor alpha receptor: cDNA cloning and demonstration that mesoderm induction establishes the lineage-specific pattern of ligand and receptor gene expression. Dev. Genet. 14, 185±193. Kaetzel, D. M., Coyne, D. W., and Fenstermaker, R. A. (1993). Transcriptional control of the platelet-derived growth factor subunit genes. Biofactors 4, 71±81. Kaetzel, D.M. Jr., Maul, R. S., Liu, B., Bonthron, D., Fenstermaker, R. A., and Coyne, D. W. (1994). Platelet-derived growth factor A-chain gene transcription is mediated by positive and negative regulatory regions in the promoter. Biochem. J. 301, 321±327. Kaetzel, D. M., Morgan, D. III., Reid, J. D. IV., and Fenstermaker, R. A. (1996). Site-directed mutagenesis of the N-linked glycosylation site in platelet-derived growth factor Bchain results in diminished intracellular retention. Biochem. Biophys. Acta 1298, 250±260. Kallio, E. A., Koskinen, P. K., Aavik, E., Buchdunger, E., and Lemstrom, K. B. (1999). Role of platelet-derived growth factor in obliterative bronchiolitis (chronic rejection) in the rat. Am. J. Respir. Crit. Care Med. 160, 1324±1332. Karlsson, L., Bondjers, C., and Betsholtz, C. (1999). Roles for PDGF-A and sonic hedgehog in development of mesenchymal components of the hair follicle. Development 126, 2611± 2621. Kartha, S., Bradham, D. M., Grotendorst, G. R., and Toback, F. G. (1988). Kidney epithelial cells express c-sis protooncogene and secrete PDGF-like protein. Am. J. Physiol. 255, F800±806. Katayose, D., Ohe, M., Yamauchi, K., Ogata, M., Shirato, K., Fujita, H., Shibahara, S., and Takishima, T. (1993). Increased expression of PDGF A- and B-chain genes in rat lungs with hypoxic pulmonary hypertension. Am. J. Physiol. 264, L100± L106. Kato, H., Osajima, A., Uezono, Y., Okazaki, M., Tsuda, Y., Tanaka, H., Oishi, Y., Izumi, F., and Nakashima, Y. (1999). Involvement of PDGF in pressure-induced mesangial cell proliferation through PKC and tyrosine kinase pathways. Am. J. Physiol. 277, F105±F112. Katoh, O., Kimura, A., Itoh, T., and Kuramoto, A. (1990). Platelet derived growth factor messenger RNA is increased in bone marrow megakaryocytes in patients with myeloproliferative disorders. Am. J. Hematol. 35, 145±150. Kawabe, T., Wen, T. C., Matsuda, S., Ishihara, K., Otsuda, H., and Sakanaka, M. (1997). Platelet-derived growth factor prevents ischemia-induced neuronal injuries in vivo. Neurosci. Res. 29, 335±343. Kawaguchi, Y., Hara, M., and Wright, T. M. (1999). Endogenous IL-1alpha from systemic sclerosis fibroblasts induces IL-6 and PDGF-A. J. Clin. Invest. 103, 1253±1260. Keck, P. J., Hauser, S. D., Krivi, G., Sanzo, K., Warren, T., Feder, J., and Connolly, D. T. (1989). Vascular permeability factor, an endothelial cell mitogen related to PDGF. Science 246, 1309±1312. Kelly, J. L., Sanchez, A., Brown, G. S., Chesterman, C. N., and Sleight, M. J. (1993). Accumulation of PDGF B and cell-binding forms of PDGF A in the extracellular matrix. J. Cell Biol. 121, 1153±1163. Kelly, K., Cochran, B. H., Stiles, C. D., and Leder, P. (1983). Cellspecific regulation of the c-myc gene by lymphocyte mitogens and platelet-derived growth factor. Cell 35, 603±610.
PDGF 783 Kenagy, R. D., and Clowes, A. W. (1995). Regulation of baboon arterial smooth muscle cell plasminogen activators by heparin and growth factors. Thromb. Res. 77, 55±61. Kenney, W. C., Haniu, M., Herman, A. C., Arakawa, T., Costigan, V. J., Lary, J., Yphantis, D. A., and Thomason, A. R. (1994). Formation of mitogenically active PDGF-B dimer does not require interchain disulfide bonds. J. Biol. Chem. 269, 12351±12359. Keutzer, J. C., and Sytkowski, A. J. (1995). Regulated production of a pleiotropic cytokine-platelet-derived growth factor ± by differentiating erythroid cells in vitro and in vivo. Proc. Natl Acad. Sci. USA 92, 4967±4971. Khachigian, L. M., Resnick, N., Gimbrone, M. A. Jr, and Collins, T. (1995). Nuclear factor-kappa B interacts functionally with the platelet-derived growth factor B-chain shear-stress response element in vascular endothelial cells exposed to fluid shear stress. J. Clin. Invest. 96, 1169±1175. Khachigian, L. M., Anderson, K. R., Halnon, N. J., Gimbrone, M. A. Jr, Resnick, N., and Collins, T. (1997). Egr-1 is activated in endothelial cells exposed to fluid shear stress and interacts with a novel shear-stress-response element in the PDGF A-chain promoter. Arterioscler. Thromb. Vasc. Biol. 17, 2280±2286. Khachigian, L. M., Santiago, F. S., Rafty, L. A., Chan, O. L., Delbridge, G. J., Bobik, A., Collins, T., and Johnson, A. C. (1999). GC factor 2 represses platelet-derived growth factor A-chain gene transcription and is itself induced by arterial injury. Circ. Res. 84, 1258±1267. Kieswetter, K., Schwartz, Z, Alderete, M., Dean, D. D., and Boyan, B. D. (1997). Platelet-derived growth factor stimulates chondrocyte proliferation but prevents endochondral maturation. Endocrine 6, 257±264. Kim, W. J., Mohan, R. R., Mohan, R. R., and Wilson, S. E. (1999). Effect of PDGF, IL-1, and BMP2/4 on corneal fibroblast chemotaxis: expression of the platelet-derived growth factor system in the cornea. Invest. Ophthalmol. Vis. Sci. 40, 1364±1372. Kirstein, M., Brett, J., Radoff, S., Ogawa, S., Stern, D., and Vlassara, H. (1990). Advanced protein glycosylation induces transendothelial human monocyte chemotaxis and secretion of platelet-derived growth factor: role in vascular disease of diabetes and aging. Proc. Natl Acad. Sci. USA 87, 9010± 9014. Klareskog, L., Gustafsson, R., Scheynius, A., and Hallgren, R. (1990). Increased expression of platelet-derived growth factor type B receptors in the skin of patients with systemic sclerosis. Arthritis Rheum. 33, 1534±1541. Kociok, N., Heppekausen, H., Schraermeyer, U., Esser, P., Thumann, G., Grisanti, S., and Heimann, K. (1998). The mRNA expression of cytokines and their receptors in cultures iris pigment epithelial cells: a comparison with retinal pigment epithelial cells. Exp. Eye Res. 67, 237±250. Koempel, J. A., Gibson, S. E., O'Grady, K., and Toriumi, D. M. (1998). The effect of platelet-derived growth factor on tracheal wound healing. Int. J. Pediatr. Otorhinolaryngol. 46, 1±8. Kohler, N., and Lipton, A. (1974). Platelets as a source of fibroblast growth-promoting activity. Exp. Cell Res. 87, 297±301. Koyama, H., Raines, E. W., Bornfeldt, K. E., Roberts, J. M., and Ross, R. (1996). Fibrillar collagen inhibits arterial smooth muscle proliferation through regulation of Cdk2 inhibitors. Cell 87, 1069±1078. Kraiss, L. W., Raines, E. W., Wilcox, J. N., Seifert, R. A., Barrett, T. B., Kirkman, T. R., Hart, C. E., BowenPope, D. F., Ross, R., and Clowes, A. W. (1993). Regional expression of the platelet-derived growth factor and its receptors
in a primate graft model of vessel wall assembly. J. Clin. Invest. 92, 338±348. Kraiss, L. W., Geary, R. L., Mattsson, E. J., Vergel, S., Au, Y. P., and Clowes, A. W. (1996). Acute reductions in blood flow and shear stress induce platelet-derived growth factor-A expression in baboon prosthetic grafts. Circ. Res. 79, 45±53. Kreysing, J., OÈstman, A., van de Poll, M., BaÈckstroÈm, G., and Heldin, C.-H. (1996). Identification of three amino acid residues in the B-chain of platelet-derived growth factor with different importance for binding to PDGF alpha- and beta-receptors. FEBS Lett. 385, 181±184. Kwon, Y. K. (1997). Expression of brain-derived neurotrophic factor mRNA stimulated by basic fibroblast growth factor and platelet-derived growth factor in rat hippocampal cell line. Mol. Cells 7, 320±325. LaFleur, D. W., Fagin, J. A., Forrester, J. S., Rubin, S. A., and Sharifi, B. G. (1994). Cloning and characterization of alternatively spliced isoforms of rat tenascin. Platelet-derived growth factor-BB markedly stimulates expression of spliced variants of tenascin mRNA in arterial smooth muscle cells. J. Biol. Chem. 269, 20757±20763. LaMarre, J., Wollenberg, G. K., Gonias, S. L., and Hayes, M. A. (1991). Cytokine binding and clearance properties of proteinase-activated alpha 2-macroglobulins. Lab. Invest. 65, 3±14. LaRochelle, W. J., May-Siroff, M., Robbins, K. C., and Aaronson, S. A. (1991). A novel mechanism regulating growth factor association with the cell surface: identification of a PDGF retention domain. Genes Dev. 5, 1191±1199. Lee, S. B., Kolquist, K. A., Nichols, K., Englert, C., Maheswaran, S., Ladanyi, M., Gerald, W. L., and Haber, D. A. (1997). The EWS-WT1 translocation product induces PDGFA in desmoplastic small round-cell tumour. Nature Genet. 17, 309±313. LeGrand, E. K. (1998). Preclinical promise of becaplermin (rhPDGF-BB) in wound healing. Am. J. Surg. 176, 48S±54S. Leon, S. P., Carroll, R. S., Dashner, K., Glowacka, D., and Black, P. M. (1994). Messenger ribonucleic acid expression of platelet-derived growth factor subunits and receptors in pituitary adenomas. J. Clin. Endocrinol. Metab. 79, 51±55. LeveÂen, P., Betsholtz, C., and Westermark, B. (1993). Negative trans-acting mechanisms controlling expression of plateletderived growth factor A and B mRNA in somatic cell hybrids. Exp. Cell Res. 207, 283±289. LeveÂen, P., Pekny, M., Gebre-Medhim, S., Swolin, B., Larsson, E., and Betsholtz, C. (1994). Mice deficient for PDGF B show renal, cardiovascular, and hematological abnormalities. Genes Dev. 8, 1875±1887. Li, S., PonteÂn, A., Aase, K., Karlsson, L., Abramsson, A., Uutela, M., BaÈckstroÈm, G., HellstroÈm, M., BostroÈm, H., Li, H., Soriano, P., Betsholtz, C., Heldin, C.-H., Alitalo, K., OÈstman, A., and Eriksson, U. (2000). PDGF-C is a new protease-activated ligand for the PDGF -receptor. Nature Cell Biol. 2, 302±309. Liang, Y., Robinson, D. F., Dennig, J., Suske, G., and Fahl, W. E. (1996). Transcriptional regulation of the SIS/PDGF-B gene in human osteosarcoma cells by the Sp family of transcription factors. J. Biol. Chem. 271, 11792±11797. Libby, P. Warner, S. J., Salomon, R. N., and Birinyi, L. K. (1988). Production of platelet-derived growth factor-like mitogen by smooth muscle cells from human atheroma. N. Engl. J. Med. 318, 1493±1498. Liechty, K. W., Nesbit, M., Herlyn, M., Radu, A., Adzick, N. S., and Crombleholme, T. M. (1999). Adenoviral-mediated overexpression of platelet-derived growth factor-B corrects ischemic impaired wound healing. J. Invest. Dermatol. 113, 375±383.
784 E. W. Raines Limanni, A., Fleming, T., Molina, R., Hufnagel, H., Cunningham, R. E., Cruess, D. F., and Sharefkin, J. B. (1988). Expression of genes for platelet-derived growth factor in adult human venous endothelium. A possible non-plateletdependent cause of intimal hyperplasia in vein grafts and perianastomotic areas of vascular prostheses. J. Vasc. Surg. 7, 10±20. Lin, X., Wang, Z., Gu, L., and Deuel, T. F. (1992). Functional analysis of the human platelet-derived growth factor A-chain promoter region. J. Biol. Chem. 267, 25614±25619. Lin, X.-H., Guo, C., Gu, L. J., and Deuel, T. F. (1993). Sitespecific methylation inhibits transcriptional activity of platelet-derived growth factor A-chain promoter. J. Biol. Chem. 268, 17334±17340. Lindahl, P., and Betsholtz, C. (1998). Not all myofibroblasts are alike: revisiting the role of PDGF-A and PDGF-B using PDGF-targeted mice. Curr. Opin. Nephrol. Hypertens. 7, 21±26. Lindahl, P., Johansson, B. R., LeveÂen, P., and Betsholtz, C. (1997a). Pericyte loss and microaneurysm formation in PDGFB-deficient mice. Science 277, 242±245. Lindahl, P., Karlsson, L., HellstroÈm, M., Gebre-Medhim, S., Willetts, K., Heath, J. K., and Betsholtz, C. (1997b). Alveogenesis failure in PDGF-A-deficient mice is coupled to lack of distal spreading of alveolar smooth muscle cell progenitors during lung development. Development 124, 3943±3953. Lindahl, P., HellstroÈm, M., Kalen, M., Karlsson, L., Pekny, M., Pekna, M., Soriano, P., and Betsholtz, C. (1998). Paracrine PDGF-B/PDGF-Rbeta signaling controls mesangial cell development in kidney glomeruli. Development 125, 3313±3322. Lindahl, P., BostroÈm, H., Karlsson, L., HellstroÈm, M., Kalen, M., and Betsholtz, C. (1999). Role of platelet-derived growth factors in angiogenesis and alveogenesis. Curr. Top. Pathol. 93, 27±33. Ling, B. N., Matsugana, H., Ma, H., and Eaton, D. C. (1995). Role of growth factors in mesangial cell ion channel regulation. Kidney Int. 48, 1158±1166. Liu, J. Y., Brass, D. M., Hoyle, G. W., and Brody, A. R. (1998). TNF-alpha receptor knockout mice are protected from the fibroproliferative effects of inhaled asbestos fibers. Am. J. Pathol. 153, 1839±1847. Liu, P., Ying, Y., Ko, Y.-G., and Anderson R. G. W. (1996). Localization of platelet-derived growth factor-stimulated phosphorylation cascade to caveolae. J. Biol. Chem. 271, 10299± 10303. Lustig, F., Hoebeke, J., Simonson, C., Ostergren-Lunden, G., Bondjers, G., Ruetchi, U., and Fager, G. (1999). Processing of PDGF gene products determines interactions with glycosaminoglycans. J. Mol. Recognit. 12, 112±120. Lynch, S. E., de Castilla, G. R., Williams, R. C., Kiritsy, C. P., Howell, T. H., Reddy, M. S., and Antoniades, H. N. (1991). The effects of short-term application of a combination of platelet-derived and insulin-like growth factors on periodontal wound healing. J. Periodontol. 62, 458±467. McKay, N. G., Hunter, D. J., Haites, N. E., and Power, D. A. (1994). Regulation of alternative splicing of the fibronectin IIICS domain by cytokines. Biochem. Biophys. Res. Commun. 199, 1005±1011. Majesky, M. W., Reidy, M. A., Bowen-Pope, D. F., Hart, C. E., Wilcox, J. N., and Schwartz, S. M. (1990a). PDGF ligand and receptor gene expression during repair of arterial injury. J. Cell Biol. 111, 2149±2158. Majesky, M. W., Daemen, M. D., and Schwartz, S. M. (1990b). Alpha 1-adrenergic stimulation of platelet-derived growth factor A-chain gene expression in rat aorta. J. Biol. Chem. 265, 1082±1088.
Malden, L. T., Chait, A., Raines, E. W., and Ross, R. (1991). The influence of oxidatively modified low density lipoproteins on expression of platelet-derived growth factor by human monocyte-derived macrophages. J. Biol. Chem. 266, 13901±13907. Maldonado, P. E., Rose, B., and Loewenstein, W. R. (1988). Growth factors modulate junctional cell-to-cell communication. J. Membr. Biol. 106, 203±210. Marmur, R., Mabie, P. C., Gokhan, S., Song, Q., Kessler, J. A., and Mehler, M. F. (1998). Isolation and developmental characterization of cerebral cortical multipotent progenitors. Dev. Biol. 204, 577±591. Marra, F., Choudhury, G. G., Pinzani, M., and Abboud, H. E. (1994). Regulation of platelet-derived growth factor secretion and gene expression in human liver fat-storing cells. Gastroenterol. 107, 1110±1117. Matsuda, M., Shikata, K., Makino, H., Sugimoto, H., Ota, K., Akiyama, K., Hirata, K., and Ota, Z. (1997). Gene expression of PDGF and PDGF receptor in various forms of glomerulonephritis. Am. J. Nephrol. 17, 25±31. Maxwell, M., Naber, S. P., Wolfe, H. J., Galanopoulos, T., Hedley-Whyte, E. T., Black, P. M., and Antoniades, H. N. (1990). Coexpression of platelet-derived growth factor (PDGF) and PDGF-receptor genes by primary human astrocytomas may contribute to their development and maintenance. J. Clin. Invest. 86, 131±140. Mehta, D., George, S. J., Jeremy, J. Y., Izzat, M. B., Southgate, K. M., Bryan, A. J., Newby, A. C., and Angelini, G. D. (1998). External stenting reduces long-term medial and neointimal thickening and platelet-derived growth factor expression in a pig model of arteriovenous bypass grafting. Nature Med. 4, 235±239. Meier, C., Parmantier, E., Brennan, A., Mirsky, R., and Jessen, K. R. (1999). Developing Schwann cells acquire the ability to survive without axons by establishing an autocrine circuit involving insulin-like growth factor, neurotrophin-3, and platelet-derived growth factor-BB. J. Neurosci. 19, 3847±3859. Mekada, A., Sasahara, M., Yamada, E., Kani, K., and Hazama, F. (1998). Platelet-derived growth factor B-chain expression in the rat retina and optic nerve: distribution and changes after transection of the optic nerve. Vision Res. 38, 3031±3039. Mendoza, A. E., Young, R., Orkin, S. H., and Collins, T. (1990). Increased platelet-derived growth factor A-chain expression in human uterine smooth muscle cells during the physiologic hypertrophy of pregnancy. Proc. Natl Acad. Sci. USA 87, 2177±2181. Mercola, M., Melton, D. A., and Stiles, C. D. (1988). Plateletderived growth factor A chain is maternally encoded in Xenopus embryos. Science 241, 1223±1225. Mercurius, K. O., and Morla, A. O. (1998). Inhibition of vascular smooth muscle cell growth by inhibition of fibronectin matrix assembly. Circ. Res. 82, 548±556. Miller, M. S. (1999). Use of topical recombinant human plateletderived growth factor-BB (becaplermin) in healing of chronic mixed arteriovenous lower extremity diabetic ulcers. J. Foot Ankle Surg. 38, 227±231. Morita, T., and Kourembanas, S. (1995). Endothelial cell expression of vasoconstrictors and growth factors is regulated by smooth muscle cell-derived carbon monoxide. J. Clin. Invest. 96, 2676±2682. Mudhar, H. S., Pollock, R. A., Wang, C., Stiles, C. D., and Richardson, W. D. (1993). PDGF and its receptors in the developing rodent retina and optic nerve. Development 118, 539±552. Muller, Y. A., Christinger, H. W., Keyt, B. A., and de Vos, A. M. (1997). The crystal structure of vascular endothelial growth
PDGF 785 factor (VEGF) refined to 1.93 A resolution: multiple copy flexibility and receptor binding. Structure 5, 1325±1338. Murray-Rust, J., McDonald, N. Q., Blundell, T. L., Hosang, M., Oefner, C., Winkler, F., and Bradshaw, R. A. (1993). Topological similarities in TGF-beta 2, PDGF-BB and NGF define a superfamily of polypeptide growth factors. Structure 1, 153±159. Nagaoka, I., Trapnell, B. C., and Crystal, R. G. (1990). Upregulation of platelet-derived growth factor-A and -B gene expression in alveolar macrophages of individuals with idiopathic pulmonary fibrosis. J. Clin. Invest. 85, 2023±2027. Nakamura, T., Ebihara, I., Nagaoka, I., Tomino, Y., and Koide, H. (1992). Activated peripheral blood mononuclear cells in IgA nephropathy express platelet-derived growth factor B-chain messenger RNA. J. Lab. Clin. Med. 120, 212±221. Nakano, T., Raines, E. W., Abraham, J. A., Wenzel, F.G. IV, Higashiyama, S., Klagsbrun, M., and Ross, R. (1993). Glucocorticoid inhibits thrombin-induced expression of plateletderived growth factor A-chain and heparin-binding EGF-like growth factor in human aortic smooth muscle cells. J. Biol. Chem. 268, 22941±22947. Nakashima, N., Umeda, F., Yamauchi, T., Ishii, H., Hisatomi, A., Nawata, H., Masuko, H., Nakayama, K., and Tatematsu, A. (1992). Platelet-derived growth factor and growth-promoting activity in the serum samples and platelets of patients with non-insulin-dependent diabetes mellitus. J. Lab. Clin. Med. 120, 78±85. Narisawa-Saito, M., Silva, A. J., Yamaguchi, T., Hayashi, T., Yamamoto, T., and Nawa, H. (1999). Growth factor-mediated Fyn signaling regulates alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptor expression in rodent neocortical neurons. Proc. Natl Acad. Sci. USA 96, 2461± 2466. Negoro, N., Kanayama, Y., Haraguchi, M., Umetani, N., Nishimura, M., Konishi, Y, Iwai, J, Okamura, M., Inoue, T., and Takeda, T. (1995). Blood pressure regulates platelet-derived growth factor A-chain gene expression in vascular smooth muscle cells in vivo. An autocrine mechanism promoting hypertensive vascular hypertrophy. J. Clin. Invest. 95, 1140±1150. Nikkhah, G., Odin, P., Smits, A., Tingstrom, A., Othberg, A., Brundin, P., Funa, K., and Lindvall, O. (1993). Platelet-derived growth factor promotes survival of rat and human mesencephalic dopaminergic neurons in culture. Exp. Brain Res. 92, 516± 523. NisteÂr, M., Libermann, T. A., Betsholtz, C., Pettersson, M., Claesson-Welsh, L., Heldin, C.-H., Schlessinger, J., and Westermark, B. (1988). Expression of messenger RNAs for platelet-derived growth factor and transforming growth factoralpha and their receptors in human malignant glioma cell lines. Cancer Res. 48, 3910±3918. Noble, M., Murray, K., Stroobant, P., Waterfield, M. D., and Riddle, P. (1988). Platelet-derived growth factor promotes division and motility and inhibits premature differentiation of the oligodendrocyte/type-2 astrocyte progenitor cell. Nature 333, 560±562. Nobuyoshi, M., Lin, X.-H., Takimoto, Y., Deuel, T. F., and Wang, Z.-Y. (1997). Transcription regulation of the PDGF A-chain gene by first intron elements. Biochem. Biophys. Res. Commun. 230, 569±572. Oefner, C., D'Arcy, A., Winkler, F. K., Eggimann, B., and Hosang, M. (1992). Crystal structure of human platelet-derived growth factor BB. EMBO J. 11, 3921±3926. Ohlsson, R., Falck, P., HellstroÈm, M., Lindahl, P., BostroÈm, H., Franklin, G., Ahrlund-Richter, L., Pollard, J., Soriano, P., and Betsholtz, C. (1999). PDGFB regulates the development of the
labyrinthine layer of the mouse fetal placenta. Dev. Biol. 212, 124±136. Ohno, M., Sasahara, M., Narumiya, S., Tanaka, N., Yamano, T., Shimada, M., and Hazama, F. (1999). Expression of plateletderived growth factor B-chain and beta-receptor in hypoxic/ ischemic encephalopathy of neonatal rats. Neuroscience 90, 643±651. Okazaki, H., Majesky, M. W., Harker, L. A., and Schwartz, S. M. (1992). Regulation of platelet-derived growth factor ligand and receptor gene expression by alpha-thrombin in vascular smooth muscle cells. Circ. Res. 71, 1285±1293. Ono, I., Gunji, H., Zhang, J. Z., Maruyama, K., and Kaneko, F. (1995). A study of cytokines in burn blister fluid related to wound healing. Burns 21, 352±355. Oquendo, P., Alberta, J., Wen, D. Z., Graycar, J. L., Derynck, R., and Stiles, C. D. (1989). The platelet-derived growth factorinducible KC gene encodes a secretory protein related to platelet alpha-granule proteins. J. Biol. Chem. 264, 4133±4137. OÈstman, A., Rall, L., Hammacher, A., Wormstead, M. A., Coit, D., Valenzuela, P, Betsholtz, C., Westermark, B., and Heldin, C.-H. (1988). Synthesis and assembly of a functionally active recombinant platelet-derived growth factor AB heterodimer. J. Biol. Chem. 263, 16202±16208. OÈstman, A., Thyberg, J., Westermark, B., and Heldin, C.-H. (1992). PDGF-AA and PDGF-BB biosynthesis: Proprotein processing in the Golgi complex and lysosomal degradation of PDGF-BB retained intracellularly. J. Cell Biol. 118, 509±519. Pantazis, P., Sariban, E., Kufe, D., and Antoniades, H. N. (1986). Induction of c-sis gene expression and synthesis of plateletderived growth factor in human myeloid leukemia cells during monocytic differentiation. Proc. Natl Acad. Sci. USA 83, 6455± 6459. Pantazis, P., Sariban, E., Bohan, C. A., Antoniades, H. N., and Kalyanaraman, V. S. (1987). Synthesis of PDGF by cultured human T cells transformed with HTLV-I and II. Oncogene 1, 285±289. Papayannopoulou, T., Raines, E., Collins, S., Nakamoto, B., Tweeddale, M., and Ross, R. (1987). Constitutive and inducible secretion of platelet-derived growth factor analogs by human leukemic cell lines coexpressing erythroid and megakaryocytic markers. J. Clin. Invest. 79, 859±866. Paulsson, Y., Hammacher, A., Heldin, C.-H., and Westermark, B. (1987). Possible positive autocrine feedback in the prereplicative phase of human fibroblasts. Nature 328, 715±717. Paulsson, Y., Austgulen, R., Hofsli, E., Heldin, C.-H., Westermark, B., and Nissen-Meyer, J. (1989). Tumor necrosis factor-induced expression of platelet-derived growth factor A-chain messenger RNA in fibroblasts. Exp. Cell Res. 180, 490±496. Pech, M., Rao, C. D., Robbins, K. C., and Aaronson, S. A. (1989). Functional identification of regulatory elements within the promoter region of platelet-derived growth factor 2. Mol. Cell Biol. 9, 396±405. Peres, R., Betsholtz, C., Westermark, B., and Heldin, C.-H. (1987). Frequent expression of growth factors for mesenchymal cells in human mammary carcinoma cell lines. Cancer Res. 47, 3425±3429. Pfeilschifter, J., Krempien, R., Naumann, A., Gronwald, R. G., Hoppe, J., and Ziegler, R. (1992). Differential effects of plateletderived growth factor isoforms on plasminogen activator activity in fetal rat osteoblasts due to isoform-specific receptor functions. Endocrinology 130, 2059±2066. Pierce, G. F., Mustoe, T. A., Altrock, B. W., Deuel, T. F., and Thomason, A. (1991). Role of platelet-derived growth factor in wound healing. J. Cell Biochem. 45, 319±326.
786 E. W. Raines Pierce, G. F., Tarpley, J. E., Tseng, J., Bready, J., Chang, D., Kenney, W. C., Rudolph, R., Robson, M. D., Vande Berg, J., Reid, P., Kaufman, S., and Farrell, C. L. (1995). Detection of platelet-derived growth factor (PDGF)-AA in actively healing human wounds treated with recombinant PDGF-BB and absence of PDGF in chronic nonhealing wounds. J. Clin. Invest. 96, 1336±1350. Pinzani, M., Milani, S., Herbst, H., DeFranco, R., Grappone, C., Gentilini, A., Caligiuri, A., Pellegrini, G., Ngo, D. V., Romanelli, R. G., and Gentilini, P. (1996). Expression of platelet-derived growth factor and its receptors in normal human liver and during active hepatic fibrogenesis. Am. J. Pathol. 148, 785±800. Plata-SalamaÂn, C. R. (1988). Food intake suppression by growth factors and platelet peptides by direct action in the central nervous system. Neurosci. Lett. 94, 161±166. Pollock, R. A., and Richardson, W. D. (1992). The alternativesplice isoforms of the PDGF A-chain differ in their ability to associate with the extracellular matrix and to bind heparin in vitro. Growth Factors 7, 267±277. Potts, J. D., Bassnett, S., Kornacker, S., and Beebe, D. C. (1994). Expression of platelet-derived growth factor receptors in the developing chicken lens. Invest. Ophthalmol. Vis. Sci. 35, 3413±3421. Pringle, N. P., and Richardson, W. D. (1993). A singularity of PDGF alpha-receptor expression in the dorsoventral axis of the neural tube may define the origin of the oligodendrocyte lineage. Development 117, 525±533. Qiu, C. X., and Ferguson, M. W. (1995). The distribution of PDGFs and PDGF-receptors during murine secondary palate development. J. Anat. 186, 17±29. Raff, M. C., Lillien, L. E., Richardson, W. D., Burne, J. F., and Noble, M. D. (1988). Platelet-derived growth factor from astrocytes drives the clock that times oligodendrocyte development in culture. Nature 333, 562±565. Rafty, L. A., and Khachigian, L. M. (1998). Zinc finger transcription factors mediate high constitutive platelet-derived growth factor-B expression in smooth muscle cells derived from aortae of newborn rats. J. Biol. Chem. 273, 5758±5764. Raines, E. W., and Ross, R. (1982). Platelet-derived growth factor. I: High yield purification and evidence for multiple forms. J. Biol. Chem. 257, 5154±5160. Raines, E. W., and Ross, R. (1992). Compartmentalization of PDGF on extracellular binding sites dependent on exon-6encoded sequences. J. Cell Biol. 116, 533±543. Raines, E. W., and Ross, R. (1993). In ``Cytokines, Vol. 5, The Biology of PDGF'' (ed B. Westermark and C. Sorg), Plateletderived growth factor in vivo pp. 74±114.S. Karger, Basel Raines, E. W., and Ross, R. (1996). Multiple growth factors are associated with lesions of atherosclerosis: specificity or redundancy? BioEssays 18, 271±282. Raines, E. W., Bowen-Pope, D. F., and Ross, R. (1984). Plasma binding proteins for platelet-derived growth factor that inhibit its binding to cell-surface receptors. Proc. Natl Acad. Sci. USA 81, 3424±3428. Raines, E. W., Dower, S. K., and Ross, R. (1989). IL-1 mitogenic activity for fibroblasts and smooth muscle cells is due to PDGFAA. Science 243, 393±396. Raines, E. W., Bowen-Pope, D. F., and Ross, R. (1990). In ``Handbook of Experimental Pharmacology: Peptide Growth Factors and Their Receptors I, Vol. 95/I'' (ed M. B. Sporn and A. B. Roberts), Platelet-derived growth factor pp. 173± 262. Springer-Verlag. New York. Raines, E. W., Lane, T. F., Iruela-Arispe, L., Ross, R., and Sage, E. H. (1992). The extracellular glycoprotein SPARC interacts with platelet-derived growth factor (PDGF)-AB
and -BB and inhibits the binding of PDGF to its receptors. Proc. Natl Acad. Sci. USA 89, 1281±1285. Raines, E. W., Koyama, H., and Carragher, N. O. (2000). The extracellular matrix dynamically regulates smooth muscle cell responsiveness to PDGF. (In press). Ramachandran, R. K., Wikramanayake, A. H., Uzman, J. A., Govindarajan, V., and Tomlinson, C. R. (1997). Disruption of gastrulation and oral-aboral ectoderm differentiation in the Lytechinus pictus embryo by a dominant/negative PDGF receptor. Development 124, 2355±2364. Rappolee, D. A., Brenner, C. A., Schultz, R., Mark, D., and Werb, Z. (1988). Developmental expression of PDGF, TGFalpha, and TGF-beta genes in preimplantation mouse embryos. Science 241, 1823±1825. Rechtoris, C., and Mazzone, T. (1995). Isoform-specific induction of the low-density lipoprotein receptor gene by platelet-derived growth factor. Am. J. Physiol. 268, C1033±C1039. Rees, R. S., Robson, M. C., Smiell, J. M., and Perry, B. H. (1999). Becaplermin gel in the treatment of pressure ulcers: a phase II randomized, double-blind, placebo-controlled study. Wound Repair Regen. 7, 141±147. Reilly, C. F., and McFall, R. C. (1991). Platelet-derived growth factor and transforming growth factor-beta regulate plasminogen activator inhibitor-1 synthesis in vascular smooth muscle cells. J. Biol. Chem. 266, 9419±9427. Reneker, L. W., and Overbeek, P. A. (1996a). Lens-specific expression of PDGF-A in transgenic mice results in retinal astrocytic hamartomas. Invest. Ophthalmol. Vis. Sci. 37, 2455±2466. Reneker, L. W., and Overbeek, P. A. (1996b). Lens specific expression of PDGF-A alters lens growth and development. Dev. Biol. 180, 554±565. Resnick, N., Collins, T., Atkinson, W., Bonthron, D. T., Dewey Jr, C. F., and Gimbrone, M. A. Jr (1993). Plateletderived growth factor B chain promoter contains a cis-acting fluid shear-stress-responsive element. Proc. Natl Acad. Sci. USA 90, 4591±4595. Reuning, U., and Bang, N. U. (1992). Regulation of the urokinase-type plasminogen activator receptor on vascular smooth muscle cells is under the control of thrombin and other mitogens. Arterioscler. Thromb. 12, 1161±1170. Reuterdahl, C., TingstroÈm, A., Terracio, L., Funa, K., Heldin, C.H., and Rubin, K. (1991). Characterization of platelet-derived growth factor beta-receptor expressing cells in the vasculature of human rheumatoid synovium. Lab. Invest. 64, 321±329. Reuterdahl, C., Sundberg, C., Rubin, K., Funa, K., and Gerdin, B. (1993). Tissue localization of beta receptors for platelet-derived growth factor and platelet-derived growth factor B chain during wound repair in humans. J. Clin. Invest. 91, 2065±2075. Rice, A. B., Moomaw, C. R., Morgan, D. L., and Bonner, J. C. (1999). Specific inhibitors of platelet-derived growth factor or epidermal growth factor receptor tyrosine kinase reduce pulmonary fibrosis in rats. Am. J. Pathol. 155, 213±221. Richardson, W. D., Pringle, N., Mosley, M. J., Westermark, B., and Dubois-Dalcq, M. (1998). A role for platelet-derived growth factor in normal gliogenesis in the central nervous system. Cell 53, 309±319. Risau, W., Drexler, H., Mironov, V., Smits, A., Siegbahn, A., Funa, K., and Heldin, C.-H. (1992). Platelet-derived growth factor is angiogenic in vivo. Growth Factors 7, 261±266. Robbins, K. C., Antoniades, H. N., Devare, S. G., Hunkapiller, M. W., and Aaronson, S. A. (1983). Structural and immunological similarities between simian sarcoma virus gene products(s) and human platelet-derived growth factor. Nature 305, 605±608.
PDGF 787 Robbins, S. G., Mixon, R. N., Wilson, D. J., Hart, C. E., Robertson, J. E., Westra, I., Planck, S. R., and Rosenbaum, J. T. (1994). Platelet-derived growth factor ligands and receptors immunolocalized in proliferative retinal diseases. Invest. Ophthalmol. Vis. Sci. 35, 3649±3663. Robson, M. C., Mustoe, T. A., and Hunt, T. K. (1998). The future of recombinant growth factors in wound healing. Am. J. Surg. 176, 80S±82S. Rodt, S. A., Ahlen, K., Berg, A., Rubin, K., and Reed, R. K. (1996). A novel physiological function for platelet-derived growth factor-BB in rat dermis. J. Physiol. (Lond.) 495, 193± 200. Rollins, B. J. (1991). JE/MCP-1: an early-response gene encodes a monocyte-specific cytokine. Cancer Cells 3, 517±524. Rollins, B. J., and Stiles, C. D. (1988). Regulation of c-myc and cfos proto-oncogene expression by animal cell growth factors. In Vitro Cell Dev. Biol. 24, 81±84. Rollins, B. J., Morrison, E. D., and Stiles, C. D. (1988). Cloning and expression of JE, a gene inducible by platelet-derived growth factor and whose product has cytokine-like properties. Proc. Natl Acad. Sci. USA 85, 3738±3742. Rorsman, F., Bywater, M., Knott, T. J., Scott, J., and Betsholtz, C. (1988). Structural characterization of the human plateletderived growth factor A-chain cDNA and gene: alternative exon usage predicts two different precursor proteins. Mol. Cell. Biol. 8, 571±577. Ross, R. (1968). The fibroblast and wound repair. Biol. Rev. 43, 51±96. Ross, R. (1999). Atherosclerosis: an inflammatory disease. N. Engl. J. Med. 340, 115±126. Ross, R., Glomset, J., Kariya, B., and Harker, L. (1974). A platelet-dependent serum factor that stimulates the proliferation of arterial smooth muscle cells in vitro. Proc. Natl Acad. Sci. USA 71, 1207±1210. Ross, R., Raines, E. W., and Bowen-Pope, D. F. (1986). The biology of platelet-derived growth factor. Cell 46, 155±169. Ross, R., Masuda, J., Raines, E. W., Gown, A. M., Katsuda, S., Sasahara, M., Malden, L. T., Masuko, H., and Sato, H. (1990). Localization of PDGF-B protein in macrophages in all phases of atherogenesis. Science 248, 1009±1012. Rubin, K., TingstroÈm, A., Hansson, G. K., Larsson, E., RoÈnnstrand, L., Klareskog, L., Claesson-Welsh, L., Heldin, C.-H., FellstroÈm, B., and Terracio, L. (1988). Induction of B-type receptors for platelet-derived growth factor in vascular inflammation: possible implications for development of vascular proliferative lesions. Lancet 1, 1353±1356. Rydziel, S., Shaikh, S., and Canalis, E. (1994). Platelet-derived growth factor-AA and -BB (PDGF-AA and -BB) enhance the synthesis of PDGF-AA in bone cell cultures. Endocrinology 134, 2541±2546. Santiago, F. S., Atkins, D. G., and Khachigian, L. M. (1999). Vascular smooth muscle cell proliferation and regrowth after mechanical injury in vitro are Egr-1/NGFI-A-dependent. Am. J. Pathol. 155, 897±905. Sariban, E., and Kufe, D. (1988). Expression of the plateletderived growth factor 1 and 2 genes in human myeloid cell lines and monocytes. Cancer Res. 48, 4498±4502. Sariban, E., Sitaras, N. M., Antoniades, H. N., Kufe, D. W., and Pantazis, P. (1988). Expression of platelet-derived growth factor (PDGF)-related transcripts and synthesis of biologically active PDGF-like proteins by human malignant epithelial cell lines. J. Clin. Invest. 82, 1157±1164. Sasahara, M., Fries, J. W. U., Raines, E. W., Gown, A. M., Westrum, L. E., Frosch, M. P., Bonthron, D. T., Ross, R., and Collins, T. (1991). PDGF B-chain in neurons of the central
nervous system, posterior pituitary, and in a transgenic model. Cell 64, 217±227. Sasahara, M., Amano, S., Sato, H., Yang, J. G., Hayase, Y., Kaneko, M., Sato, I., Suzaki, M., and Hazama, F. (1998). Normal developing rat brain expresses a platelet-derived growth factor B chain (c-sis) mRNA truncated at the 50 end. Oncogene 16, 1571±1578. Sasaki, K.. Oomura, Y., Suzuki, K., Muto, T., Hanai, K., Tooyama, I., Kimura, H., and Yanaihara, N. (1991). Effects of fibroblast growth factors and platelet-derived growth factor on food intake in rats. Brain Res. Bull. 27, 327±332. Sato, N., Beitz, J. G., Kato, J., Yamamoto, M., Clark, J. W., Calabresi, P., Raymond, A., and Frackelton Jr, A. R. (1993). Platelet-derived growth factor indirectly stimulates angiogenesis in vitro. Am. J. Pathol. 142, 1119±1130. Sawdey, M. S., and Loskutoff, D. J. (1991). Regulation of murine type 1 plasminogen activator inhibitor gene expression in vivo. Tissue specificity and induction by lipopolysaccharide, tumor necrosis factor-alpha, and transforming growth factor-beta. J. Clin. Invest. 88, 1346±1353. Scarpati, E. M., and DiCorleto, P. E. (1996). Identification of a thrombin response element in the human platelet-derived growth factor B-chain (c-sis) promoter. J. Biol. Chem. 271, 3025±3032. Schilling, D., Reid, J.D. IV, Hujer, A., Morgan, D., Demoll, E., Bummer, P., Fenstermaker, R. A., and Kaetzel, D. M. (1998). Loop III region of platelet-derived growth factor (PDGF) Bchain mediates binding to PDGF receptors and heparin. Biochem. J. 333, 637±644. SchoÈnherr, E., JaÈrvelaÈinen, H. T., Sandell, L. J., and Wight, T. N. (1991). Effects of platelet-derived growth factor and transforming growth factor-beta 1 on the synthesis of a large versican-like chondroitin sulfate proteoglycan by arterial smooth muscle cells. J. Biol. Chem. 266, 17640±17647. SchoÈnherr, E., JaÈrvelaÈinen, H. T., Kinsella, M. G., Sandell, L. J., and Wight, T. N. (1993). Platelet-derived growth factor and transforming growth factor-beta 1 differentially affect the synthesis of biglycan and decorin by monkey arterial smooth muscle cells. Arterioscler. Thromb. 13, 1026±1036. Seifert, R. A., Schwartz, S. M., and Bowen-Pope, D. F. (1984). Developmentally regulated production of platelet-derived growth factor-like molecules. Nature 311, 669±671. Seifert, R. A., Hart, C. E., Phillips, P. E., Forstrom, J. W., Ross, R., Murray, M. J., and Bowen-Pope, D. F. (1989). Two different subunits associate to create isoform-specific plateletderived growth factor receptors. J. Biol. Chem. 264, 8771±8778. Seifert, R. A., Alpers, C. E., and Bowen-Pope, D. F. (1998). Expression of platelet-derived growth factor and its receptors in the developing and adult mouse kidney. Kidney Int. 54, 731±746. Sejersen, T., Betsholtz, C., Sjolund, M., Heldin, C.-H., Westermark, B., and Thyberg, J. (1986). Rat skeletal myoblasts and arterial smooth muscle cells express the gene for the A chain but not the gene for the B chain (c-sis) of platelet-derived growth factor (PDGF) and produce a PDGF-like protein. Proc. Natl Acad. Sci. USA 83, 6844±6848. Sella, O., Gerlitz, G., Le, S. Y., and Elroy-Stein, O. (1999). Differentiation-induced internal translation of c-sis mRNA: analysis of the cis elements and their differentiation-linked binding to the hnRNP C protein. Mol. Cell Biol. 19, 5429±5440. Seppa, H., Grotendorst, G., Seppa, S., Schiffmann, E., and Martin, G. R. (1982). Platelet-derived growth factor is chemotactic for fibroblasts. J. Cell Biol. 92, 584±588. Shaddy, R. E., Hansen, J. C., and Cowley, C. G. (1992). Effects of T cells on platelet-derived growth factor-like protein secretion from endothelial cells. J. Heart Lung Transplant. 11, 48±56.
788 E. W. Raines Shaddy, R. E., Hammond, E. H., and Yowell, R. L. (1996). Immunohistochemical analysis of platelet-derived growth factor and basic fibroblast growth factor in cardiac biopsy and autopsy specimens of heart transplant patients. Am. J. Cardiol. 77, 1210±1215. Shimokado, K., Raines, E. W., Madtes, D. K., Barrett, T. B., Benditt, E. P., and Ross, R. (1985). A significant part of macrophage-derived growth factor consists of at least two forms of PDGF. Cell 43, 277±286. Shultz, P. J., DiCorleto, P. E., Silver, B. J., and Abboud, H. E. (1988). Mesangial cells express PDGF mRNAs and proliferate in response to PDGF. Am. J. Physiol. 255, F674±684. Shultz, P. J., Knauss, T. C., MeneÂ, P., and Abboud, H. E. (1989). Mitogenic signals for thrombin in mesangial cells: regulation of phospholipase C and PDGF genes. Am. J. Physiol. 257, F366± F374. Shure, D., Senior, R. M., Griffin, G. L., and Deuel, T. F. (1992). PDGF AA homodimers are potent chemoattractants for fibroblasts and neutrophils, and for monocytes activated by lymphocytes or cytokines. Biochem. Biophys. Res. Commun. 186, 1510±1514. Sihvola, R., Koskinen, P., Myllarniemi, M., Loubtchenkov, M., HaÈyry, P., Buchdunger, E., and Lemstrom, K. (1999). Prevention of cardiac allograft arteriosclerosis by protein tyrosine kinase inhibitor selective for platelet-derived growth factor receptor. Circulation 99, 2295±2301. Silberstein, F. C., De Simone, R., Levi, G., and Aloisi, F. (1996). Cytokine-regulated expression of platelet-derived growth factor gene and protein in cultured human astrocytes. J. Neurochem. 66, 1409±1417. Silver, B. J., Jaffer, F. E., and Abboud, H. E. (1989). Plateletderived growth factor synthesis in mesangial cells: induction by multiple peptide mitogens. Proc. Natl Acad. Sci. USA 86, 1056±1060. Silverman, E. S., Khachigian, L. M., Lindner, V., Williams, A. J., and Collins, T. (1997). Inducible PDGF A-chain transcription in smooth muscle cells is mediated by Egr-1 displacement of Sp1 and Sp3. Am. J. Physiol. 273, H1415±H1426. Simon, M. P., Pedeutour, F., Sirvent, N., Grosgeorge, J., Minoletti, F., Coindre, J. M., Terrier-Lacombe, M. J., Mandahl, N., Craver, R. D., Blin, N. Sozzi, G., TurcCarel, C., O'Brien, K. P., Kedra, D., Fransson, I., Guildbaud, C., and Dumanski, J. P. (1997). Deregulation of the platelet-derived growth factor B-chain gene via fusion with collagen gene COL1A1 in dermatofibrosarcoma protuberans and giant-cell fibroblastoma. Nature Genet. 15, 95±98. Singh, J. P., Chaikin, M. A., and Stiles, C. D. (1982). Phylogenetic analysis of platelet-derived growth factor by radio-receptor assay. J. Cell Biol. 95, 667±671. Sitaras, N. M., Sariban, E., Pantazis, P., Zetter, B., and Antoniades, H. N. (1987). Human iliac artery endothelial cells express both genes encoding the chains of platelet-derived growth factor (PDGF) and synthesize PDGF-like mitogen. J. Cell. Physiol. 132, 376±380. Smiell, J. M. (1998). Clinical safety of becaplermin (rhPDGF-BB) gel.Becaplermin Studies Group. Am. J. Surg. 176, 68S±73S. Smits, A., and Funa, K. (1998). Platelet-derived growth factor (PDGF) in primary brain tumours of neuroglial origin. Histol. Histopathol. 13, 511±520. Smits, A., Hermansson, M., NisteÂr, M., Karnushina, I., Heldin, C.-H., Westermark, B., and Funa, K. (1989). Rat brain capillary endothelial cells express functional PDGF B-type receptors. Growth Factors 2, 1±8. Smits, A., Kato, M., Westermark, B., NisteÂr, M., Heldin, C.-H., and Funa, K. (1991). Neurotrophic activity of platelet-derived
growth factor (PDGF): rat neuronal cells possess functional PDGF beta-type receptors and respond to PDGF. Proc. Natl Acad. Sci. USA 88, 8159±8163. Soriano, P. (1994). Abnormal kidney development and hematological disorders in PDGF beta-receptor mutant mice. Genes Dev. 8, 1888±1896. Soriano, P. (1997). The PDGF alpha receptor is required for neural crest cell development and for normal patterning of the somites. Development 124, 2691±2700. Sprugel, K. H., McPherson, J. M., Clowes, A. W., and Ross, R. (1987). Effects of growth factors in vivo. I: Cell ingrowth into porous subcutaneous chambers. Am. J. Pathol. 129, 601±613. Stenman, G., Rorsman, F., and Betsholtz, C. (1988). Sublocalization of the human PDGF A-chain gene to chromosome 7, band q11.23, by in situ hybridization. Exp. Cell Res. 178, 180±184. Sturzl, M., Roth, W. K., Brockmeyer, N. H., Zietz, C., Speiser, B., and Hofschneider, P. H. (1992). Expression of platelet-derived growth factor and its receptor in AIDS-related Kaposi sarcoma in vivo suggests paracrine and autocrine mechanisms of tumor maintenance. Proc. Natl Acad. Sci. USA 89, 7046±7050. Swan, D. C., McBride, O. W., Robbins, K. C., Keithley, D. A., Reddy, E. P., and Aaronson, S. A. (1982). Chromosomal mapping of the simian sarcoma virus onc gene analogue in human cells. Proc. Natl Acad. Sci. USA 79, 4691±4695. Takase, H., Oemar, B. S., Pech, M., and Luscher, T. F. (1999). Platelet-derived growth factor-induced vasodilation in mesenteric resistance arteries by nitric oxide: blunted response in spontaneous hypertension. J. Cardiovasc. Pharmacol. 33, 223± 228. Tamura, M., and Iwamoto, Y. (1989). The effect of plateletderived growth factor on phagocytosis of cultured human trabecular cells. Exp. Eye Res. 48, 761±770. Tang, W. W., Ulich, T. R., Lacey, D. L., Hill, D. C., Qi, M., Kaufman, S. A., Van, G. Y., Tarpley, J. E., and Yee, J. S. (1996). Platelet-derived growth factor-BB induces renal tubulointerstitial myofibroblast formation and tubulointerstitial fibrosis. Am. J. Pathol. 148, 1169±1180. Taniguchi, Y., Yorioka, N., Oda, H., and Yamakido, M. (1996). Platelet-derived growth factor, interleukin (IL)-1 beta, IL-6R and tumor necrosis factor-alpha in IgA nephropathy. An immunohistochemical study. Nephron 74, 652±660. Taniuchi, K., Yamada, Y., Nonomura, A., and Takehara, K. (1997). Immunohistochemical analysis of platelet-derived growth factor and its receptors in fibrohistiocytic tumors. J. Cutan. Pathol. 24, 393±397. Tanizawa, S., Ueda, M., van der Loos, C. M., van der Wal, A. C., and Becker, A. E. (1996). Expression of platelet-derived growth factor B chain and beta receptor in human coronary arteries after percutaneous transluminal coronary angioplasty: an immunohistochemical study. Heart 75, 549±556. Taylor, R. N., and Williams, L. T. (1988). Developmental expression of platelet-derived growth factor and its receptor in the human placenta. Mol. Endocrinol. 2, 627±632. Tiesman, J., and Hart, C. E. (1993). Identification of a soluble receptor for platelet-derived growth factor in cell-conditioned medium and human plasma. J. Biol. Chem. 268, 9621± 9628. Tischer, E., Gospodarowicz, D., Mitchell, R., Silva, M., Schilling, J., Lau, K., Crisp, T., Fiddes, J. C., and Abraham, J. A. (1989). Vascular endothelial growth factor: a new member of the platelet-derived growth factor gene family. Biochem. Biophys. Res. Commun. 165, 1998±1206. Trink, B., Wang, G., Shahar, M., Meydan, N., and Roifman, C. M. (1995). Functional platelet-derived growth
PDGF 789 factor-beta (PDGF-beta) receptor expressed on early B-lineage precursor cells. Clin. Exp. Immunol. 102, 417±424. Tsai, L. H., White, L., Raines, E., Ross, R., Smith, R. G., Cushley, W., and Ozanne, B. (1994). Expression of plateletderived growth factor and its receptors by two pre-B acute lymphocytic leukemia cell lines. Blood 83, 51±55. Uebelhoer, M., Bewig, B., Kreipe, H., Nowak, D., Magnussen, H., and Barth, J. (1995). Modulation of fibroblast activity in histiocytosis X by platelet-derived growth factor. Chest 107, 701±705. Valenzuela, C. F., Kazlauskas, A., Brozowski, S. J., Weiner, J. L., Demali, K. A., McDonald, B. J., Moss, S. J., Dunwiddie, T. V., and Harris, R. A. (1995). Platelet-derived growth factor receptor is a novel modulator of type A gamma-aminobutyric acidgated ion channels. Mol. Pharmacol. 48, 1099±1107. Valenzuela, C. F., Xiong, Z., MacDonld, J. F., Weiner, J. L., Frazier, C. J., Dunwiddie, T. V., Kazlauskas, A., Whiting, P. J., and Harris, R. A. (1996). Platelet-derived growth factor induces a long-term inhibition of N-methyl-D-aspartate receptor function. J. Biol. Chem. 271, 16151±16159. Valenzuela, C. F., Kazlauskas, A., and Weiner, J. L. (1997). Roles of platelet-derived growth factor in the developing and mature nervous systems. Brain Res. Brain Res. Rev. 24, 77±89. Van Zoelen, E. J., van de Ven, W. J., Franssen, H. J., van Oostwaard, T. M., van der Saag, P. T., Heldin, C.-H., and de Laat, S. W. (1985). Neuroblastoma cells express c-sis and produce a transforming growth factor antigenically related to the platelet-derived growth factor. Mol. Cell Biol. 5, 2289± 2297. Vassbotn, F. S., OÈstman, A., Langeland, N., Holmsen, H., Westermark, B., Heldin, C.-H., and NisteÂr, M. (1994). Activated platelet-derived growth factor autocrine pathway drives the transformed phenotype of a human glioblastoma cell line. J. Cell Physiol. 158, 381±389. Vesaluoma, M., Teppo, A. M., Gronhagen-Riska, C., and Tervo, T. (1997). Platelet-derived growth factor-BB (PDGFBB) in tear fluid: a potential modulator of corneal wound healing following photorefractive keratectomy. Curr. Eye Res. 16, 825±831. Vogt, P. M., Lehnhardt, M., Wagner, D., Jansen, V., Krieg, M., and Steinau, H. U. (1998). Determination of endogenous growth factors in human wound fluid: temporal presence and profiles of secretion. Plast. Reconstr. Surg. 102, 117±123. Wang, C., and Stiles, C. D. (1993). Regulation of platelet-derived growth factor A messenger RNA translation in differentiating F9 teratocarcinoma cells. Cell Growth Differ. 4, 871±877. Wang, D., Huang, H. J., Kazlauskas, A., and Cavenee, W. K. (1999). Induction of vascular endothelial growth factor expression in endothelial cells by platelet-derived growth factor through the activation of phosphatidylinositol 3-kinase. Cancer Res. 59, 1464±1472. Wang, L., Hayashi, H., and Ebina, Y. (1999). Transient effect of platelet-derived growth factor on GLUT4 translocation in 3T3L1 adipocytes. J. Biol. Chem. 274, 19246±19253. Wang, Z. Y., and Deuel, T. F. (1992). An S1 nuclease-sensitive homopurine/homopyrimidine domain in the PDGF A-chain promoter contains a novel binding site for the growth factorinducible protein EGR-1. Biochem. Biophys. Res. Commun. 188, 433±439. Wang, Z. Y., Madden, S. L., Deuel, T. F., and Rauscher, F. J. III (1992). The Wilms' tumor gene product, WT1, represses transcription of the platelet-derived growth factor A-chain gene. J. Biol. Chem. 267, 21999±22002. Wang, Z.-Y., Masaharu, N., Qiu, Q.-Q., Takimoto, Y., and Deuel, T. F. (1994). An S1 nuclease-sensitive region in the
first intron of human platelet-derived growth factor A-chain gene contains a negatively acting cell type-specific regulatory element. Nucleic Acids Res. 22, 457±464. Watanabe, S., Wang, X. E., Hirose, M., Oide, H., Kitamura, T., Miwa, H., Miyazaki, A., and Sato, N. (1996). Platelet-derived growth factor accelerates gastric epithelial restoration in a rabbit cultured cell model. Gastroenterology 110, 775±779. Weiler-Guettler, H., Aird, W. C., Rayburn, H., Husain, M., and Rosenberg, R. D. (1996). Developmentally regulated gene expression of thrombomodulin in postimplantation mouse embryos. Development 122, 2271±2278. Weima, S. M., van Rooijen, M. A., Mummery, C. L., Feijen, A., Kruijer, W., de Laat, S. W., and van Zoelen, E. J. (1988). Differentially regulated production of platelet-derived growth factor and of transforming growth factor beta by a human teratocarcinoma cell line. Differentiation 38, 203±210. Westermark, B., and Wasteson, AÊ. (1976). A platelet factor stimulating human normal glial cells. Exp. Cell Res. 98, 170±174. Westermark, B., Johnsson, A., Paulsson, Y., Betsholtz, C., Heldin, C.-H., Herlyn, M., Rodeck, U., and Koprowski, H. (1986). Human melanoma cell lines of primary and metastatic origin express the genes encoding the chains of platelet-derived growth factor (PDGF) and produce a PDGF-like growth factor. Proc. Natl Acad. Sci. USA 83, 7197±7200. Wickenhauser, C., Hillienhof, A., Jungheim, K., Lorenzen, J., Ruskowski, H., Hansmann, M. L., Thiele, J., and Fisher, R. (1995). Detection and quantification of transforming growth factor beta (TGF-beta) and platelet-derived growth factor (PDGF) release by normal human megakaryocytes. Leukemia 9, 310±315. Wilcox, J. N., Smith, K. M., Williams, L. T., Schwartz, S. M., and Gordon, D. (1988). Platelet-derived growth factor mRNA detection in human atherosclerotic plaques by in situ hybridization. J. Clin. Invest. 82, 1134±1143. Williams, B. P., Park, J. K., Alberta, J. A., Muhlebach, S. G., Hwang, G. Y., Roberts, T. M., and Stiles, C. D. (1997). A PDGF-regulated immediate early gene response initiates neuronal differentiation in ventricular zone progenitor cells. Neuron 18, 553±562. Williams, L. T., Antoniades, H. N., and Goetzl, E. J. (1983). Platelet-derived growth factor stimulates mouse 3T3 cell mitogenesis and leukocyte chemotaxis through different structural determinants. J. Clin. Invest. 72, 1759±1763. Wilson, E., Vives, F., Collins, T., and Ives, H. E. (1998). Strainresponsive regions in the platelet-derived growth factor-A gene promoter. Hypertension 31, 170±175. Wolswijk, G., and Noble, M. (1992). Cooperation between PDGF and FGF converts slowly dividing O-2A adult progenitor cells to rapidly dividing cells with characteristics of O-2A perinatal progenitor cells. J. Cell Biol. 118, 889±900. Wong, J., Rauhoft, C., Dilley, R. J., Agrotis, A., Jennings, G. L., and Bobik, A. (1997). Angiotensin-converting enzyme inhibition abolishes medial smooth muscle PDGF-AB biosynthesis and attenuates cell proliferation in injured carotid arteries: relationships to neointima formation. Circulation 96, 1631±1640. Wu, S. M., Patel, D. D., and Pizzo, S. V. (1998). Oxidized alpha2macroglobulin (alpha2M) differentially regulates receptor binding by cytokines/growth factors: implications for tissue injury and repair mechanisms in inflammation. J. Immunol. 161, 4356± 4365. Yablonka-Reuveni, Z. (1995). Development and postnatal regulation of adult myoblasts. Microsc. Res. Tech. 30, 366±380. Yablonka-Reuveni, Z., and Seifert, R. A. (1993). Proliferation of chicken myoblasts is regulated by specific isoforms of plateletderived growth factor: evidence for differences between
790 E. W. Raines myoblasts from mid and late stages of embryogenesis. Dev. Biol. 156, 307±318. Yamaguchi, M., Gallati, H., Baur, W., Cruess, D. F., and Sharefkin, J. B. (1994). Both lisinopril and verapamil reduced platelet-derived growth factor-A chain mRNA levels in human saphenous vein endothelial cells stimulated by thrombin. Surgery 115, 495±502. Yan, X. Q., Brady, G., and Iscove, N. N. (1993). Platelet-derived growth factor (PDGF) activates primitive hematopoietic precursors (pre-CFCmulti) by up-regulating IL-1 in PDGF receptor-expressing macrophages. J. Immunol. 150, 2440±2448. Yan, X. Q., Brady, G., and Iscove, N. N. (1994). Overexpression of PDGF-B in murine hematopoietic cells induces a lethal myeloproliferative syndrome in vivo. Oncogene 9, 163±173. Yang, M., Chesterman, C. N., and Chong, B. H. (1995). Recombinant PDGF enhances megakaryocytopoiesis in vitro. Br. J. Haematol. 91, 285±289. Yeh, H. J., Ruit, K. G., Wang, Y. X., Parks, W. C., Snider, W. D., and Duel, T. F. (1991). PDGF A-chain gene in expressed by mammalian neurons during development and in maturity. Cell 64, 209±216. Yi, E. S., Lee, H., Yin, S., Piguet, P., Sarosi, I., Kaufmann, S., Tarpley, J., Wang, N. S., and Ulich, T. R. (1996). Plateletderived growth factor causes pulmonary cell proliferation and collagen deposition in vivo. Am. J. Pathol. 149, 539±548. Yoshida, K., Kuniyasu, H., Yasui, W., Kitadai, Y., Toge, T., and Tahara, E. (1993). Expression of growth factors and their receptors in human esophageal carcinomas: regulation of expression by epidermal growth factor and transforming growth factor alpha. J. Cancer Res. Clin. Oncol. 119, 401±407.
Young, R. M., Mendoza, A. E., Collins, T., and Orkin, S. H. (1990). Alternatively spliced platelet-derived growth factor Achain transcripts are not tumor specific but encode normal cellular proteins. Mol. Cell Biol. 10, 6051±6054. Zerwes, H. G., and Risau, W. (1987). Polarized secretion of a platelet-derived growth factor-like chemotactic factor by endothelial cells in vitro. J. Cell Biol. 105, 2037±2041. Zhang, F. X., Pan, W., and Hutchins, J. B. (1995). Phosphorylation of F1F0 ATPase delta-subunit is regulated by plateletderived growth factor in mouse cortical neurons in vitro. J. Neurochem. 65, 2812±2815. Zhao, X. M., Frist, W. H., Yeoh, T. K., and Miller, G. G. (1995). Confirmation of alternatively spliced platelet-derived growth factor-A chain and correlation with expression of PDGF receptor-alpha in human cardiac allografts. Transplantation 59, 605±611.
LICENSED PRODUCTS PDGF-BB is the active ingredient in REGRANEX (becaplermin) which has been approved in the United States for the treatment of lower extremity diabetic neuropathic ulcers that extend into the subcutaneous tissue or beyond and have an adequate blood supply.