+ All Categories
Home > Documents > Urokinase Regulates Vitronectin Binding by Controlling Urokinase Receptor Oligomerization

Urokinase Regulates Vitronectin Binding by Controlling Urokinase Receptor Oligomerization

Date post: 22-Nov-2023
Category:
Upload: independent
View: 0 times
Download: 0 times
Share this document with a friend
10
Urokinase Regulates Vitronectin Binding by Controlling Urokinase Receptor Oligomerization* Received for publication, December 10, 2001, and in revised form, May 17, 2002 Published, JBC Papers in Press, May 28, 2002, DOI 10.1074/jbc.M111736200 Nicolai Sidenius‡, Annapaola Andolfo, Riccardo Fesce§, and Francesco Blasi From the Molecular Genetics Unit and §Center of Theoretical Biology, DIBIT, Universita ` Vita-Salute San Raffaele, 20132 Milan, Italy Adhesion of monocytes to the extracellular matrix is mediated by a direct high affinity interaction between cell-surface urokinase-type plasminogen activator (uPA) receptor (uPAR) and the extracellular matrix pro- tein vitronectin. We demonstrate a tight connection be- tween uPA-regulated uPAR oligomerization and high affinity binding to immobilized vitronectin. We find that binding of soluble uPAR (suPAR) to immobilized vitronectin is strictly ligand-dependent with a linear relationship between the observed binding and the con- centration of ligand added. Nevertheless, a comparison of experimentally obtained binding curves to those gen- erated using a simple equilibrium model suggests that the high affinity vitronectin-binding pro-uPAsuPAR complex contains two molecules of suPAR. In co-immu- noprecipitation experiments, using different epitope- tagged suPAR molecules, suPAR/suPAR co-immunopre- cipitation displayed a similar uPA dose dependence as that observed for vitronectin binding, demonstrating that the high affinity vitronectin-binding complex in- deed contains oligomeric suPAR. Structurally, the krin- gle domain of uPA was found to be critical for the for- mation of the vitronectin-binding competent complex because the amino-terminal fragment, but not the growth factor-like domain, behaved as a full-length uPA. Our data represent the first demonstration of func- tional, ligand-induced uPAR oligomerization having ex- tensive implications for glycosylphosphatidylinositol- anchored receptors in general, and for the biology of the uPA/uPAR system in particular. Cell migration and invasion are important processes in many patho/physiological conditions such as tumor invasion, angio- genesis, and inflammation. Plasminogen activators, their in- hibitors, and their cell-surface receptor(s) play central roles in these processes by regulating extracellular proteolysis, cell ad- hesion, and signal transduction. In tissues, extracellular pro- teolysis is controlled by the production of plasmin that is gen- erated by plasminogen activators, mainly urokinase (uPA) 1 (1), which binds to a specific membrane receptor, uPAR. Fully processed human uPAR is a 45–55-kDa glycoprotein linked to the outer membrane leaflet by a glycosylphosphati- dylinositol lipid anchor (2). The protein is composed of three homologous domains with a disulfide bonding pattern charac- teristic of the uPAR/Ly-6 superfamily (3). Besides providing the cells with the means to perform di- rected extracellular matrix degradation, binding of uPA to uPAR has profound effects on cell adhesion, migration, and proliferation (4 – 6). Although binding to uPAR is always re- quired, these latter processes are often independent of the proteolytic activity of uPA, strongly suggesting that other pro- tein interactions are involved. Indeed, several data indicate that a conformational change in uPAR is capable of profoundly modifying its biological proper- ties. First, it has been shown that uPA binding to uPAR causes the appearance of novel binding sites for vitronectin (Vn) (7–10), thrombospondin (8), uPAR-associated protein (11), and the dis- appearance of binding sites for the 2 -macroglobulin receptor (8). Furthermore, uPA-induced chemotaxis (12–16) can be mimicked by proteolytic cleavage of uPAR that generates uPAR fragments that act as potent inducers of chemotaxis in cells lacking endog- enous uPAR (16, 17). However, although extensive evidence has been presented that uPAR entertains complex interactions with other proteins, very little is known about how these interactions are regulated at the molecular level. In this paper we have addressed the possibility that uPA binding influences uPAR oligomerization and that uPAR oli- gomerization is a major determinant for its interaction with other proteins. As a paradigm, we have employed the ability of uPAR to bind Vn, a function that has been shown previously to induce cell adhesion (7, 10) and to change gene expression during the differentiation of human myeloid U937 cells (18). EXPERIMENTAL PROCEDURES Materials—A soluble variant of uPAR (residues 1–277) was ex- pressed and purified from culture supernatants of transfected CHO cells as described previously (16). SuPAR/FLAG was purified from transiently transfected COS-7 cells as described previously (17). Pro- uPA, purified from eukaryotic cell culture supernatants, was a kind gift from Dr. Jack Henkin (Abbott). The amino-terminal fragments (ATF) of uPA and LMW-uPA were obtained from American Diagnostica. The growth factor-like domain of uPA (GFD, amino acids 1– 48) was a kind gift from Dr. Steve Rosenberg. The exact molarity of suPAR and pro- uPA was established by amino acid analysis (Research Consortium Inc.) Urea-purified vitronectin was purchased from Promega. The mono- clonal antibody R2 was kindly provided by Dr. Gunilla Høyer-Hansen (Finsen Laboratories, Copenhagen, Denmark). The monoclonal anti- body M2 and the pNPP substrate were obtained from Sigma. Secondary antibodies were from Amersham Biosciences (peroxidase-conjugated) and Dako (alkaline phosphatase-conjugated). Vn Binding Assays—Binding assays were performed in 96-well plates (NUNC Maxisorb) coated with Vn (0.1 ml/well, 1 g/ml in 0.05 M phosphate buffer, pH 9.6) overnight at 4 °C and blocked with 0.15 ml of 2% BSA in PBS for 1 h. All subsequent incubations were performed with reagents diluted in dilution buffer (PBS containing 1% BSA) at * The costs of publication of this article were defrayed in part by the payment of page charges. This article must therefore be hereby marked advertisement” in accordance with 18 U.S.C. Section 1734 solely to indicate this fact. ‡ To whom correspondence should be addressed: Molecular Genetics Unit DIBIT, Universita ` Vita-Salute San Raffaele Via Olgettina 58, 20132 Milan, Italy. Tel.: 39-02-26434758; Fax: 39-02-26434844; E-mail: [email protected]. 1 The abbreviations used are: uPA, urokinase-type plasminogen acti- vator; uPAR, uPA receptor; suPAR, soluble uPAR; Vn, vitronectin; BSA, bovine serum albumin; PBS, phosphate-buffered saline; GFD, growth factor-like domain; ATF, amino-terminal fragment; CD, catalytic do- main; KD, kringle domain. THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 277, No. 31, Issue of August 2, pp. 27982–27990, 2002 © 2002 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A. This paper is available on line at http://www.jbc.org 27982 by guest on February 17, 2016 http://www.jbc.org/ Downloaded from
Transcript

Urokinase Regulates Vitronectin Binding by Controlling UrokinaseReceptor Oligomerization*

Received for publication, December 10, 2001, and in revised form, May 17, 2002Published, JBC Papers in Press, May 28, 2002, DOI 10.1074/jbc.M111736200

Nicolai Sidenius‡, Annapaola Andolfo, Riccardo Fesce§, and Francesco Blasi

From the Molecular Genetics Unit and §Center of Theoretical Biology, DIBIT, Universita Vita-SaluteSan Raffaele, 20132 Milan, Italy

Adhesion of monocytes to the extracellular matrix ismediated by a direct high affinity interaction betweencell-surface urokinase-type plasminogen activator(uPA) receptor (uPAR) and the extracellular matrix pro-tein vitronectin. We demonstrate a tight connection be-tween uPA-regulated uPAR oligomerization and highaffinity binding to immobilized vitronectin. We find thatbinding of soluble uPAR (suPAR) to immobilizedvitronectin is strictly ligand-dependent with a linearrelationship between the observed binding and the con-centration of ligand added. Nevertheless, a comparisonof experimentally obtained binding curves to those gen-erated using a simple equilibrium model suggests thatthe high affinity vitronectin-binding pro-uPA�suPARcomplex contains two molecules of suPAR. In co-immu-noprecipitation experiments, using different epitope-tagged suPAR molecules, suPAR/suPAR co-immunopre-cipitation displayed a similar uPA dose dependence asthat observed for vitronectin binding, demonstratingthat the high affinity vitronectin-binding complex in-deed contains oligomeric suPAR. Structurally, the krin-gle domain of uPA was found to be critical for the for-mation of the vitronectin-binding competent complexbecause the amino-terminal fragment, but not thegrowth factor-like domain, behaved as a full-length uPA.Our data represent the first demonstration of func-tional, ligand-induced uPAR oligomerization having ex-tensive implications for glycosylphosphatidylinositol-anchored receptors in general, and for the biology of theuPA/uPAR system in particular.

Cell migration and invasion are important processes in manypatho/physiological conditions such as tumor invasion, angio-genesis, and inflammation. Plasminogen activators, their in-hibitors, and their cell-surface receptor(s) play central roles inthese processes by regulating extracellular proteolysis, cell ad-hesion, and signal transduction. In tissues, extracellular pro-teolysis is controlled by the production of plasmin that is gen-erated by plasminogen activators, mainly urokinase (uPA)1 (1),which binds to a specific membrane receptor, uPAR.

Fully processed human uPAR is a 45–55-kDa glycoproteinlinked to the outer membrane leaflet by a glycosylphosphati-dylinositol lipid anchor (2). The protein is composed of threehomologous domains with a disulfide bonding pattern charac-teristic of the uPAR/Ly-6 superfamily (3).

Besides providing the cells with the means to perform di-rected extracellular matrix degradation, binding of uPA touPAR has profound effects on cell adhesion, migration, andproliferation (4–6). Although binding to uPAR is always re-quired, these latter processes are often independent of theproteolytic activity of uPA, strongly suggesting that other pro-tein interactions are involved.

Indeed, several data indicate that a conformational changein uPAR is capable of profoundly modifying its biological proper-ties. First, it has been shown that uPA binding to uPAR causesthe appearance of novel binding sites for vitronectin (Vn) (7–10),thrombospondin (8), uPAR-associated protein (11), and the dis-appearance of binding sites for the �2-macroglobulin receptor (8).Furthermore, uPA-induced chemotaxis (12–16) can be mimickedby proteolytic cleavage of uPAR that generates uPAR fragmentsthat act as potent inducers of chemotaxis in cells lacking endog-enous uPAR (16, 17). However, although extensive evidence hasbeen presented that uPAR entertains complex interactions withother proteins, very little is known about how these interactionsare regulated at the molecular level.

In this paper we have addressed the possibility that uPAbinding influences uPAR oligomerization and that uPAR oli-gomerization is a major determinant for its interaction withother proteins. As a paradigm, we have employed the ability ofuPAR to bind Vn, a function that has been shown previously toinduce cell adhesion (7, 10) and to change gene expressionduring the differentiation of human myeloid U937 cells (18).

EXPERIMENTAL PROCEDURES

Materials—A soluble variant of uPAR (residues 1–277) was ex-pressed and purified from culture supernatants of transfected CHOcells as described previously (16). SuPAR/FLAG was purified fromtransiently transfected COS-7 cells as described previously (17). Pro-uPA, purified from eukaryotic cell culture supernatants, was a kind giftfrom Dr. Jack Henkin (Abbott). The amino-terminal fragments (ATF) ofuPA and LMW-uPA were obtained from American Diagnostica. Thegrowth factor-like domain of uPA (GFD, amino acids 1–48) was a kindgift from Dr. Steve Rosenberg. The exact molarity of suPAR and pro-uPA was established by amino acid analysis (Research Consortium Inc.)Urea-purified vitronectin was purchased from Promega. The mono-clonal antibody R2 was kindly provided by Dr. Gunilla Høyer-Hansen(Finsen Laboratories, Copenhagen, Denmark). The monoclonal anti-body M2 and the pNPP substrate were obtained from Sigma. Secondaryantibodies were from Amersham Biosciences (peroxidase-conjugated)and Dako (alkaline phosphatase-conjugated).

Vn Binding Assays—Binding assays were performed in 96-wellplates (NUNC Maxisorb) coated with Vn (0.1 ml/well, 1 �g/ml in 0.05 M

phosphate buffer, pH 9.6) overnight at 4 °C and blocked with 0.15 ml of2% BSA in PBS for 1 h. All subsequent incubations were performedwith reagents diluted in dilution buffer (PBS containing 1% BSA) at

* The costs of publication of this article were defrayed in part by thepayment of page charges. This article must therefore be hereby marked“advertisement” in accordance with 18 U.S.C. Section 1734 solely toindicate this fact.

‡ To whom correspondence should be addressed: Molecular GeneticsUnit DIBIT, Universita Vita-Salute San Raffaele Via Olgettina 58,20132 Milan, Italy. Tel.: 39-02-26434758; Fax: 39-02-26434844; E-mail:[email protected].

1 The abbreviations used are: uPA, urokinase-type plasminogen acti-vator; uPAR, uPA receptor; suPAR, soluble uPAR; Vn, vitronectin; BSA,bovine serum albumin; PBS, phosphate-buffered saline; GFD, growthfactor-like domain; ATF, amino-terminal fragment; CD, catalytic do-main; KD, kringle domain.

THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 277, No. 31, Issue of August 2, pp. 27982–27990, 2002© 2002 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A.

This paper is available on line at http://www.jbc.org27982

by guest on February 17, 2016http://w

ww

.jbc.org/D

ownloaded from

room temperature on an orbital shaker. Initially, wells were incubatedwith suPAR and other reagents diluted to the indicated concentrations for1 h at room temperature (Note: as described under “Results,” the way thatreagents are mixed is important in determining the level of Vn bindingobtained. To minimize possible experimental variability caused by thisfact, the reagents were always mixed from 2-fold concentrated stocks(when 2 reagents were to be tested together) or 3-fold concentrated stocks(when 3 reagents were to be tested together), etc. and allowed to sit for�30 min at room temperature before being transferred to the Vn-coatedwells.) After extensive washing with PBS containing 0.1% Tween 20(PBS-T), the wells were probed for bound suPAR or suPAR/FLAG (asindicated) using monoclonal antibodies (2 �g/ml) specific for the relevantmolecule (R2 and M2, respectively) for 30 min at room temperature. Afterwashing, the wells were probed for bound antibody by incubation for 30min at room temperature with an alkaline phosphatase-conjugated rabbitanti-mouse antibody diluted 1:1000. After another round of washing,bound alkaline phosphatase activity was assayed using the chromogenicsubstrate p-nitrophenyl phosphate and quantified by measuring the ab-sorbance of samples at 405 nm in an ELISA plate reader.

Mathematical Fitting—The binding data from the experimentsshown in Fig. 1 were fitted by minimizing square errors to a simpleequilibrium model: {(i) L � R 7 LR; (ii) RL � R 7 LR2}, where L andR represent pro-uPA (ligand) and suPAR (receptor), respectively, andtotal suPAR binding to Vn is represented by B � a[LR2] � b[LR]. Inorder to fit the data it was necessary to allow for a �25% error in theestimates of the relative concentrations of pro-uPA and suPAR, perhapsbeing due to loss of the reagent because of marginal unspecific bindingto the substrate, or to other causes. For calculations of the predictedcurves presented in Fig. 1, a concentration correction factor of 0.76 forpro-uPA was used. Computations were performed in the Mathlab soft-ware environment (Matworks, Natik, MA) on a personal computer.

Immuoprecipitation and Immunoblotting Analysis—For immunopre-cipitation analysis the indicated concentrations of suPAR, suPAR/FLAG, pro-uPA, ATF, or GFD were mixed in dilution buffer and incu-bated for 1 h at room temperature. SuPAR/FLAG was thenimmunoprecipitated by addition of 20 �l of M2-agarose beads (Sigma)and incubation under gentle agitation for 1 h. After extensive washingof the beads with PBS-T, bound proteins were eluted by boiling innon-reducing SDS-PAGE sample buffer, size-fractionated by 10% SDS-PAGE, and transferred to polyvinylidene difluoride membranes bysemi-dry electroblotting (19). After blocking with 5% non-fat dry milk inPBS-T, Western blots were developed by sequential incubations withbiotinylated R2 and horseradish peroxidase-conjugated streptavidin,followed by chemiluminescent detection.

Gel Filtration Analysis—suPAR, pro-uPA, and suPAR�pro-uPA com-plexes were analyzed by size-exclusion chromatography using a Super-dexTM 200 PC 3.2/30 column (Amersham Biosciences). A sample volume of20 �l was loaded onto the column that was developed with PBS containing0.5 M NaCl using a flow rate of 60 �l/min. The elution profile was recordedby measuring the absorbance of the eluted sample at 280 nm. The columnwas calibrated by measuring the retention times of several standardsubstances, calculating their corresponding Kav values, and plotting theirKav values versus the logarithm of their molecular weights. The equationused to calculate Kav was Kav � (Ve � Vo)/(Vt � Vo), where Ve is the elutionvolume of the protein, Vo the column void volume (� elution volume ofblue dextran 2000), and Vt the total bed volume (2.4 ml). The apparentmolecular weights of pro-uPA, suPAR, and the 1:1 complex were deter-mined by linear regression. The calibration molecules, their molecularweights, and their retention times (mean of two determinations � S.D.)are as follows: thyroglobulin, 669 kDa, RT � 17.55 � 0.08 min; ferritin,440 kDa, 20.07 � 0.02 min; catalase, 232 kDa, 22.41 � 0.01 min; aldolase,158 kDa, 22.83 � 0.04 min; BSA, 67 kDa, 24.55 � 0.02 min; ovalbumin, 43kDa, 26.00 � 0.04 min; chymotrypsinogen A, 25 kDa, 28.75 � 0.01 min;ribonuclease A, 13.7 kDa, 28.75 � 0.01 min; and blue dextran 2000,�2000 kDa, 14.79 � 0.06 min.

RESULTS

Biphasic Effect of Pro-uPA on suPAR Binding to Vn—Tostudy the mechanism of uPAR binding to Vn, we have exploiteda simple in vitro binding assay in which Vn-coated plasticsurfaces are incubated with suPAR in the presence or absenceof the reagents to be tested. After removal of unbound reagents,bound suPAR was quantitated by sequential incubations withreagent-specific primary antibodies, secondary enzyme-conju-gated antibodies, and finally a colorimetric substrate (see “Ex-perimental Procedures”).

In this assay, suPAR displays specific, high affinity, andligand-dependent binding to Vn (Fig. 1A), indicating that thehigh affinity Vn-binding form of uPAR is a complex betweenuPAR and pro-uPA (7, 10). However, although pro-uPA wasclearly required for suPAR binding to Vn, it was also stronglyinhibitory when present in excess of suPAR. This suggests thatthe mechanism of suPAR binding to immobilized Vn was morecomplex than a simple reaction in which one molecule of suPARbinds one molecule of pro-uPA forming a heterodimeric highaffinity Vn-binding complex. To exclude the possibility thatthis unusual dose dependence of binding was an artifact causedby our detection system, we repeated the experiments usingdifferent antibodies to detect bound suPAR (a polyclonal rabbitantibody and two different mouse monoclonal antibodies), aswell as using 125I-radiolabeled suPAR in which case no second-ary reagent was required. However, independent of the methodof detection, qualitatively identical binding curves were ob-served in all cases (results not shown), demonstrating that thebehavior of the binding curves indeed reflects the ligand de-pendence of suPAR binding to immobilized Vn.

To address the ligand dependence of suPAR binding to Vn, wefurther analyzed the extent of suPAR binding to immobilized Vnand its dependence on pro-uPA concentration (Fig. 1B). No bind-ing of suPAR was observed in the absence of low concentrationsof pro-uPA. With increasing pro-uPA, we found a linear increasein suPAR binding, the extent of which was dependent only on thepro-uPA concentration (in the initial tract the curves superim-pose for all concentrations of suPAR). Binding reached a maxi-mum at a pro-uPA concentration close to one-half of the suPARconcentration and then declined at higher pro-uPAs. This sug-gests the existence of an optimal pro-uPA:suPAR stoichiometryfor Vn binding, possibly 1:2. At pro-uPA concentrations abovethose of suPAR, the curves settle to a plateau close to one-third ofthe maximum, suggesting the formation of complex with 1:1stoichiometry, less efficient in binding Vn.

Curves of suPAR binding to Vn for increasing concentrationsof suPAR (Fig. 1C) rise linearly at suPAR concentrations up toone-half pro-uPA concentration (and superimpose at all con-centrations of pro-uPA). At higher suPAR concentrations theslopes increase (and more so for higher pro-uPA concentra-tions) until the curves rather abruptly plateau at suPAR con-centrations about twice those of pro-uPA. This again points toan optimal 1:2 pro-uPA:suPAR stoichiometry for Vn binding.The region of linear increase in Vn binding (pro-uPA/suPAR�1/2) and the more than linear increase between stoichiometry2:1 and 1:2 again point to lower binding efficiency for com-plexes with 1:1 stoichiometry.

Qualitatively, the experimental data thus suggest that su-PAR binding to Vn may be explained by the binding of twostoichiometrically different pro-uPA�suPAR complexes. A highaffinity 1:2 complex forms when suPAR is in excess of pro-uPA,and a low affinity 1:1 complex preferentially forms when pro-uPA is in excess of suPAR. To address mathematically thevalidity of this mechanism, a simple equilibrium model was fitto the experimental data shown in Reactions 1 and 2,

KD�i�L�R4OOO3LR

REACTION 1

KD�ii�LR�R4OOO3LR2

REACTION 2

where L and R represent pro-uPA and suPAR, respectively,and total suPAR binding to Vn is represented by B � a[LR2] �b[LR]. The effectiveness ratio (b/a) represents the relative Vn

Ligand-induced Urokinase Receptor Oligomerization 27983

by guest on February 17, 2016http://w

ww

.jbc.org/D

ownloaded from

binding activity of the LR and LR2 complexes.The curves predicted by this model were slightly sensitive to

the KD values for Reactions 1 and 2, and best fits were obtainedfor low values of these parameters, KD(ii) �10�10 and KD(i)even lower (some 50-fold). This suggests that binding under theapplied experimental conditions might constitute an almostirreversible reaction, and an equilibrium model might not bethe best way to describe the system. However, most aspects ofthe experimental binding data represented in Fig. 1 were fullyand quantitatively accounted for by this model. Experimentalestimates are available in the literature for the affinity of theuPA/uPAR interaction (KD(i) 10�10 M (20–24)). Therefore, fitswere recalculated by constraining KD(i) to this value (Fig. 1, Band C, dotted lines). The resulting estimate for KD(ii) was 0.77nM, and an effectiveness ratio (b/a) of 0.061 was estimated forbinding of LR and LR2 complexes to Vn. As can be seen in Fig.1, crossing of the curves in B and the decrease to plateau in Aare well predicted.

Although the above binding model closely predicts the bind-ing curves, it should be noted that it does not take into accountthe likely presence of different multimeric forms of Vn in thebinding assay. In fact, it is possible that the high affinitypro-uPA�suPAR�Vn complex may have a 1:2:2 stoichiometryand the low affinity complex an 1:1:1 stoichiometry, etc. How-ever, binding experiments using different preparations of Vn(native and denatured) resulted in identical binding curves(data not shown) suggesting that after adsorption to plastic Vnexposes the same suPAR binding epitopes independently of itsoriginal state of oligomerization.

Pro-uPA has also been reported previously (25) to interactdirectly with Vn, and we therefore addressed the alternativepossibility that pro-uPA in some way competes for an overlap-ping binding site on Vn. However, for several reasons this couldnot explain the inhibitory effect. First, the inhibitory effect ofpro-uPA (Fig. 1) occurs at concentrations well below its appar-ent Km value for Vn (97 nM (25)). Second, the minimal concen-tration of pro-uPA required to observe inhibition was not fixedbut correlated with the concentration of suPAR (Fig. 1). Third,preincubation of immobilized Vn with pro-uPA failed to pre-vent the concomitant binding of pro-uPA�suPAR complexes(Fig. 2A, columns F–I). Another possibility is that pro-uPAcauses the release of pro-uPA�suPAR complexes from Vn. Thismight, for example, occur by proteolytic cleavage of suPAR bytrace amounts of active two-chain uPA present in the pro-uPApreparation. However, even this did not explain the inhibitoryeffect as pro-uPA�suPAR complexes bound to Vn were resistantto release by an excess of free pro-uPA (Fig. 2A, columns J–M).

Taken together the data suggest that suPAR binding to Vn isinhibited by elevated pro-uPA due to a shift from binding-competent to less effective pro-uPA�suPAR complexes for bind-ing to Vn. To test this possibility we performed binding exper-iments using pre-formed pro-uPA�suPAR complexes containingoptimal amounts of pro-uPA and suPAR (Fig. 2B). After a 1-hincubation, these pre-formed complexes were supplementedwith increasing concentrations of pro-uPA, incubated for an-other hour, and then assayed for Vn binding activity (Fig. 2B,columns F–H). For comparison, we analyzed the reaction hav-

ground absorbance observed in BSA-coated wells (see A) and is pre-sented in arbitrary units (% of the maximal binding observed in theexperiment). Individual data points of duplicate determinations areshown. C, is same as B but with a 2-fold dilution curve of suPAR in theabsence (triangles) or presence of 4.3 nM pro-uPA (circles), 8.7 nM

pro-uPA (diamonds), or 17.3 nM pro-uPA (squares). Dotted curves in Band C represent the binding curves predicted from a simple equilibriummodel: {(i) L � R 7 RL, (ii) RL � R 7 R2L}, where R and L representsuPAR and pro-uPA, respectively, and suPAR binding to Vn is repre-sented by B � a[R2L] � b[RL] (see text).

FIG. 1. Ligand dependence of suPAR binding to immobilizedVn. A, 96-well plates were coated with Vn or BSA overnight at 4 °C.Plates were washed, and the remaining binding sites were blockedusing blocking buffer (2% BSA in PBS) for 1 h at room temperature.Wells were subsequently incubated with 20.7 nM suPAR and increasingconcentrations of pro-uPA (both diluted in PBS containing 1% BSA) for1 h at room temperature. After extensive washing, bound suPAR wasquantified by sequential incubations with an anti-uPAR monoclonalantibody (R2), a secondary alkaline phosphatase conjugated anti-mouseantibody, and finally the chromogenic substrate pNPP. Data pointsrepresent the mean absorbance of triplicate samples, and the error barsindicate the S.E. of the mean. B, wells were incubated with a 2-folddilution curve of pro-uPA in the presence of 5.2 nM suPAR (triangles),10.3 nM suPAR (circles), 20.7 nM suPAR (diamonds), or 41.3 nM suPAR(squares). Vn-bound suPAR was calculated by subtracting the back-

Ligand-induced Urokinase Receptor Oligomerization27984

by guest on February 17, 2016http://w

ww

.jbc.org/D

ownloaded from

FIG. 2. Resistance of the high affinity Vn binding pro-uPA�suPAR complex to excess pro-uPA. suPAR and pro-uPA were mixed andincubated with immobilized Vn in the concentrations and sequences indicated above the graph columns, and the Vn-bound suPAR was quantifiedas described in the legend to Fig. 1. A, columns A–E, wells were incubated for 1 h with samples containing 20 nM suPAR and increasingconcentrations of pro-uPA as indicated; columns F–I, wells were first incubated for 1 h with concentrations of pro-uPA as indicated, washed, andthen incubated for 1 h with 20 nM suPAR and 10 nM pro-uPA; columns J–M, Vn-coated wells were first incubated with dilution buffer containing20 nM suPAR and 10 nM pro-uPA, washed, and then incubated with increasing concentrations of pro-uPA as indicated. B, equal volumes of dilutionbuffer containing the indicated concentrations of suPAR and pro-uPA were mixed and incubated for 1 h in the absence of Vn. Another volume ofdilution buffer with pro-uPA, as indicated, or without pro-uPA (columns A–E) was then added, and the samples were left for 1 h in the absenceof Vn. Finally, the samples were transferred to Vn-coated wells and assayed as above. (Note that this mixing scheme generates three pairs ofsamples (C/F, D/G, and E/H) having identical final concentrations of suPAR and pro-uPA, but a different mixing sequence.) Specific binding ispresented as the percentage of the maximal binding observed in the experiment, and values represent the mean (� S.D.) of triplicate determi-nations. Similar data were observed in four independent experiments.

Ligand-induced Urokinase Receptor Oligomerization 27985

by guest on February 17, 2016http://w

ww

.jbc.org/D

ownloaded from

ing the same final concentrations of both components, butprepared without the two-step addition of pro-uPA (Fig. 2B,columns B–E). Indeed, the capacity of excess pro-uPA to inhibitsuPAR binding to Vn was strongly reduced when the pro-uPA�suPAR complexes were allowed to form before the additionof the excess pro-uPA.

These data demonstrate that what determines the level ofsuPAR binding to Vn is the ratio between suPAR and pro-uPAand not the absolute concentration of the two proteins. Conse-quently, the pro-uPA�suPAR complexes that bind with highaffinity to Vn cannot be simple 1:1 complexes.

Urokinase Regulates High Affinity suPAR Binding to Vn byControlling suPAR Oligomerization—A fundamental predic-tion of the model used to explain the binding data is theexistence of complexes between pro-uPA and two (or more)molecules of suPAR.

To address this possibility directly, we performed co-immu-noprecipitation experiments using two different epitope-taggedsuPAR preparations (Fig. 3). We had previously constructed,expressed, and purified a suPAR variant in which a shortpeptide epitope had been engineered onto Pro-274 of uPARgenerating a recognition epitope for the monoclonal anti-FLAGantibody M2 (10, 17). The addition of this epitope involved theremoval of the three carboxyl-terminal amino acids (Asp-Leu-Asp, amino acids 275–277) present on “wild-type” suPAR (ami-no acids 1–277), and resulted in the complete destruction of therecognition epitope for the monoclonal anti-uPAR antibody R2.In contrast to the differential recognition by monoclonal anti-bodies, these two suPAR variants (from here on termed suPARand suPAR/FLAG) display indistinguishable pro-uPA and Vnbinding (Fig. 3 and results not shown). We mixed the two formsof suPAR and immunoprecipitated the mixture with the M2anti-FLAG antibody. When the immunoprecipitate was blottedwith R2 antibody, R2 did not recognize any protein in the M2precipitate (Fig. 3A, lane 7). However, in the presence of pro-uPA, suPAR was readily identified in the M2 immunoprecipi-tate (Fig. 3A, lane 8). The specificity of the co-immunoprecipi-tation procedure was verified by immunoprecipitation ofsamples in which one, two, or all three reactants had beenomitted (Fig. 3A, lanes 1–6). In these samples no R2 reactivematerial was observed. We next compared the pro-uPA dosedependence of suPAR co-immunoprecipitation (Fig. 3B) and Vnbinding (Fig. 3C). In these experiments an exact correlationbetween suPAR oligomerization, as evidenced by co-immuno-precipitation, and Vn binding was observed, demonstratingthat the generation of the high affinity Vn-binding complexinvolves pro-uPA-induced suPAR oligomerization.

As the suPAR:suPAR co-immunoprecipitation experimentsdemonstrated the existence of higher order suPAR�pro-uPAcomplexes in the absence of Vn, we next sought to identifythese complexes by size exclusion chromatography. To this endwe subjected purified suPAR, pro-uPA, and mixtures of the twoproteins to analytical gel filtration on a SuperdexTM 200 col-umn (Fig. 4A). Analyzed individually, both suPAR (dottedcurves) and pro-uPA (gray curves) filtered as single peaks withhighly reproducible retention times (24.80 � 0.07 (n � 3) and27.67 � 0.04 (n � 3) min, respectively). When mixtures of thesuPAR and pro-uPA were analyzed (black curves), a single newpeak (retention time, 22.91 � 0.02 min (n � 5)), correspondingto the 1:1 pro-uPA�suPAR complex was observed. No othercomplexes could be identified. Furthermore, no apparent dif-ference in the retention time of the pro-uPA�suPAR complexcould be observed when the complex was formed in the pres-ence of excess pro-uPA (22.91 � 0.01 min (n � 2)) or in thepresence of excess suPAR (22.92 � 0.01 min (n � 2)). Toexclude that the failure to detect ternary complexes by gel

filtration was caused by a lack of resolution, the column per-formance was validated using mixtures of reference proteins(Fig. 4B). In these experiments an excellent resolution wasobserved with a linear correlation (correlation coefficient 0.99)between the Kav and the log-transformed molecular weights ofthe reference proteins (see “Experimental Procedures”). Basedon the calibrations curve, the relative molecular masses ofpro-uPA, suPAR, and the 1:1 complex were calculated to be28.1 � 0.6, 75.2 � 1.1, and 144.2 � 2.2 kDa, respectively. Itthus appears that in contrast to pro-uPA, which behaves as avery compact molecule, both suPAR and the 1:1 complex mayhave an extended (non-spherical) shape in solution. Within thepresented series of gel filtrations, the maximal difference inretention time between independent runs of the same protein(11 different proteins) was 0.13 min. A reduction in the reten-tion time for the 1:1 pro-uPA�suPAR complex of twice this size(0.26 min) translated into a molecular mass increase of 13.4kDa (9.2%). We therefore concluded that if the ternary complexhad formed stoichiometrically, we would have observed it evenif its molecular size was only marginally different from that ofthe 1:1 complex.

The gel filtrations presented in Fig. 4 were all performed inthe presence of 0.5 M NaCl to prevent interactions between thesolid phase and pro-uPA which was otherwise observed atphysiological salt concentrations (data not shown). This is un-likely to explain the failure to detect higher order complexes assuPAR binding to Vn was unaffected under these conditions(data not shown). Although the use of physiological salt con-centrations did not allow a quantitative recovery of injectedpro-uPA, both suPAR and pro-uPA�suPAR complexes filterednormally under these conditions (not shown). However, evenunder these conditions no material eluting prior to the 1:1complex was ever observed (not shown). The gel filtration ex-periments thus indicated that either the higher order pro-uPA�suPAR complexes are only marginally larger that the 1:1complex or not abundant enough for detection using thismethod.

In any case the gel filtration experiments allowed us tocompare the relative abundance of free pro-uPA, suPAR, andpro-uPA�suPAR complexes, with the Vn binding activity. Tothis end, aliquots of the samples applied to the gel filtrationcolumn were diluted and assayed for their Vn binding activity(Fig. 4C). Samples containing excess suPAR (filtrations I andII) displayed approximately twice the Vn binding activity ofsamples containing equimolar concentrations of the two pro-teins (filtration III) or an excess of pro-uPA (filtrations IV andV). Indeed, the chromatograms of samples displaying a high Vnbinding activity (those in filtrations I and II) were associatedwith the presence of “free” suPAR (the left “shoulder” of thepeak corresponding to the 1:1 complex) supporting the impor-tance of unoccupied suPAR for efficient Vn binding.

The Kringle Domain of uPA Is Required to Induce uPARBinding to Vn—It has been demonstrated (9, 10) that the highaffinity interactions of uPA and Vn with uPAR require intactuPAR. We therefore focused our attention on the structuralrequirements to the ligand. Like uPAR, uPA is a modularprotein composed of three domains as follows: a small amino-terminal growth factor-like domain (GFD, uPA amino acids1–48); a central kringle domain (KD, uPA amino acids 49–135); and a carboxyl-terminal catalytic domain (CD, LMW-uPA). Although the catalytic activity of uPA is confined to theCD, the entire uPAR binding activity has been mapped to theGFD that binds to uPAR with an affinity indistinguishablefrom that of intact pro-uPA, uPA, or ATF (a proteolytic uPAfragment composed of the GFD and the KD) (23, 26, 27). Toinvestigate the structural basis for uPA promotion of suPAR

Ligand-induced Urokinase Receptor Oligomerization27986

by guest on February 17, 2016http://w

ww

.jbc.org/D

ownloaded from

oligomerization and Vn binding, we tested the capacity of dif-ferent uPA derivatives to promote suPAR binding to Vn (Fig.5A). Whereas ATF promoted suPAR binding to Vn with thesame efficiency and concentration dependence as pro-uPA, bothGFD and CD failed to do so efficiently. The failure of the GFDto promote suPAR binding to Vn was not caused by a failure tobind suPAR as both the GFD and ATF competed equally wellfor suPAR binding to pro-uPA (Fig. 5B). The weak but consis-tently observed stimulatory effect of the CD observed in Vn-binding assays and in pro-uPA competition experiments islikely to be explained by low levels of contaminating uPAand/or ATF in the commercial CD preparation used in thisstudy and was not addressed further.

As the GFD is required and sufficient for high affinity inter-action with uPAR, we concluded that receptor binding was notsufficient to promote Vn binding. In particular, the data dem-onstrate that regions within the KD are also required.

DISCUSSION

Based on the data presented in this paper we conclude thatpro-uPA controls high affinity suPAR binding to Vn by regu-lating suPAR oligomerization. Direct evidence for this hypoth-esis comes from the particular pro-uPA concentration depend-ence of suPAR binding to Vn and from pro-uPA-dependentsuPAR:suPAR co-immunoprecipitation.

Based on these data we propose a simple mechanism for theformation of the oligomeric pro-uPA�suPAR complex (Fig. 6).The first step in this reaction is the formation of the het-erodimeric pro-uPA�suPAR complex and the second the associ-ation of an unoccupied receptor molecule to form the highaffinity Vn-binding ternary pro-uPA�(suPAR)2 complex. Basedon the pro-uPA dependence of suPAR binding to Vn, we wereable to demonstrate that this ternary complex is a stable entitywhen bound to Vn (Kd � 0.77 nM, assuming a Kd � 0.1 nM for

FIG. 3. Pro-uPA regulates suPARbinding to Vn by regulating suPARoligomerization. A, purified suPAR (20nM), suPAR/FLAG (20 nM), and pro-uPA(20 nM) were mixed as indicated and in-cubated for 1 h at room temperature.SuPAR/FLAG was immunoprecipitatedusing the suPAR/FLAG-specific M2 anti-body, and the bound material was ana-lyzed by Western blotting using the su-PAR-specific monoclonal antibody R2. B,purified suPAR (20 nM), suPAR/FLAG (20nM), and the indicated concentrations ofpro-uPA were mixed and immunoprecipi-tated as described for A. C, aliquots of thereactions analyzed by co-immunoprecipi-tation in B were transferred to Vn-coatedwells and incubated for 1 h. After wash-ing, bound suPAR (squares) and suPAR/FLAG (circles) were quantified using thesuPAR-specific antibody R2 and the su-PAR/FLAG-specific antibody M2, and sec-ondary reagents as described above.

Ligand-induced Urokinase Receptor Oligomerization 27987

by guest on February 17, 2016http://w

ww

.jbc.org/D

ownloaded from

the pro-uPA-suPAR interaction).Ternary complex(es) also form in the absence of Vn, as evi-

denced by suPAR/suPAR co-immunoprecipitation (Fig. 3). Eventhough the Vn-binding curves do not allow us calculate thestability of the ternary complex in the absence of Vn, thebinding experiment presented in Fig. 2B strongly suggests thatin the absence of Vn these complexes are also stable. However,despite its predicted stability, we have not been able to identifythis complex by size exclusion chromatography (Fig. 4). How

can this apparent discrepancy be explained? Although it cannotbe excluded that the ternary complex has dimensions very closeto that of the 1:1 complex, preventing its identification by gelfiltration, the most likely explanation is that the complex ismuch less abundant than the 1:1 complex that forms quanti-tatively (Fig. 4). In support of this is the fact that even thoughsuPAR interaction with Vn is apparently of considerable affin-ity (specific binding is detectable down to sub-nanomolar lev-els), only a small fraction (estimated to be less than 5%, not

FIG. 4. Gel filtration analysis of pro-uPA�suPAR complexes. A, suPAR and pro-uPA were mixed at the indicated final concentrations andsubjected to size exclusion chromatography on a SuperdexTM 200 column, and the eluate was monitored at 280 nm. For comparison, thechromatograms obtained with the different ratios of suPAR and pro-uPA (black curves) have been overlaid with those obtained with suPAR alone(10 �M, dotted curve) and pro-uPA alone (10 �M, gray curve). B, mixtures of reference proteins with known molecular weights were subjected to gelfiltration (upper panel), and their Kav values were calculated. When the Kav values were plotted against the molecular weights (lower panel), alinear correlation (correlation coefficient � 0.99) between Kav and the log transformed molecular weights was observed underscoring the resolvingpower of the column. C, a fraction of the samples analyzed by gel filtration in A was diluted 1000-fold and assayed for Vn binding activity asdescribed in the legend to Fig. 1. The series of chromatograms shown in A represent one experiment which was repeated once with similar results.The Vn binding data shown in B represent the mean (� S.D.) of the value obtained in the two experiments.

Ligand-induced Urokinase Receptor Oligomerization27988

by guest on February 17, 2016http://w

ww

.jbc.org/D

ownloaded from

shown) of the added suPAR actually binds even at optimalconcentrations of pro-uPA. The pro-uPA and suPAR utilized inthis study have both been expressed and purified from mam-malian cell cultures and thus display the typical molecularweight heterogeneity inherent to these proteins. It is thuspossible that only a particular pattern(s) of glycosylation,and/or other type of post-translational modification(s), will al-low the formation of ternary complexes.

Although the data do not allow us to formally demonstratethe exact composition of the oligomeric pro-uPA�suPAR com-plex, it does suggest that it is a heterotrimeric complex com-posed of one ligand (pro-uPA) molecule and two receptor (su-PAR) molecules. First, the fact that we observe suPAR�suPARco-immunoprecipitation directly demonstrates that the com-plex contains at least two molecules of suPAR. Second, co-immunoprecipitation and Vn binding occur most efficientlywhen a 1:2 molar ratio between concentrations of ligand andreceptor are present, suggesting that the “active” complex con-tains a corresponding stoichiometry of the two molecules.

The predicted high affinity interaction between a het-

erodimeric pro-uPA�suPAR complex and an unoccupied suPARmolecule when bound to Vn suggests that the pro-uPA�suPAR2

complex may be stabilized by its interaction with Vn. In thiscontext the oligomeric state of Vn may be particularly impor-tant. Thermodynamically, higher order pro-uPA�suPAR�Vncomplexes, for example with a 1:2:2 stoichiometry, are expectedto be more stable than a corresponding 1:1:1 complex. In fact, aselective binding of uPAR to multimeric Vn would make senseas cells in this way could maintain the capacity to interact withthe matrix form of Vn even if surrounded by the high concen-trations of native Vn present in the circulation. Future inves-tigations on the role of Vn oligomerization in its interactionwith uPAR are highly warranted.

Implied in our model is the existence of homophilic uPAR/uPAR interaction(s) and/or divalent pro-uPA uPAR binding.Although both pro-uPA and suPAR are believed to be mono-meric proteins with monovalent interactions with each other,increasing evidence is accumulating that this may not alwaysbe true. Although the mammalian cell expressed suPAR uti-lized in this study behaves as a single defined molecular specieswhen analyzed by gel filtration (Fig. 4), it was recently shownthat when suPAR is expressed and purified from insect cellcultures it may exist as dimeric and/or multimeric aggregates(28). The aggregation of suPAR observed by Shliom et al. (28)may represent a natural homophilicity of suPAR, which isaugmented by the reduced glycosylation and/or the artificialamino-terminal amino acids present in those recombinant su-PAR variants. In support of the existence of direct homophilicuPAR/uPAR interaction(s), it was recently shown that uPAR-derived peptides may interact directly with uPAR and affect itsinteraction with both uPA and Vn (29).

We have documented that, besides being required, pro-uPAcauses a strong and consistent inhibitory effect on suPAR bind-ing to Vn and on suPAR/suPAR co-immunoprecipitation whenpresent in excess over suPAR (Figs. 1 and 3). Although themechanism of this inhibition is unclear, it evidently occursduring the formation of the oligomeric complex and not after ithas formed (Fig. 2B). The most likely mechanism is that excesspro-uPA causes a very rapid saturation of suPAR, preventingthe subsequent association between heterodimeric pro-uPA�suPAR complexes and unoccupied suPAR. Interestingly,the fact that once formed, this high affinity Vn-binding active

FIG. 5. The kringle domain of uPA is required for efficient suPAR binding to Vn. A, Vn-coated wells were incubated with 10 nM suPARand increasing concentrations of pro-uPA (squares), ATF (diamonds), GFD (circles), or LMW-uPA (triangles). After washing, bound suPAR wasquantified as described above. B, pro-uPA-coated wells were incubated with 10 nM suPAR in combination with increasing concentrations of pro-uPA(squares), ATF (diamonds), GFD (circles), and LMW-uPA (triangles). After washing, the amount of bound suPAR was determined as describedabove.

FIG. 6. Proposed molecular mechanism of uPAR binding to Vnin vitro. Proposed mechanism for the formation of high affinity and lowaffinity Vn-binding complexes between suPAR and pro-uPA is shown.In the presence of excess suPAR the presence of pro-uPA will lead to theformation of a high affinity heterotrimeric pro-uPA�(suPAR)2 complexby the association of heterodimeric pro-uPA�suPAR complexes and freesuPAR on immobilized Vn. Although this high affinity pro-uPA�suPAR2complex may form in the absence Vn, it is likely to be stabilized by itsinteraction with Vn. In the presence of excess pro-uPA, the absence offree suPAR will prevent the formation of the high affinity pro-uPA�suPAR2 complex.

Ligand-induced Urokinase Receptor Oligomerization 27989

by guest on February 17, 2016http://w

ww

.jbc.org/D

ownloaded from

complex is resistant to excess pro-uPA, suggests that it may failto bind another pro-uPA molecule.

Why has ligand-induced uPAR oligomerization escaped dis-covery until now? The key observation leading us to the dis-covery of ligand-induced dimerization of suPAR was the pecu-liar uPA dose dependence of suPAR binding to immobilized Vn(Fig. 1). However, although data on the uPAR/Vn interactionhas been presented by several groups, similar binding curveshave never been described before. The reason for this apparentdiscrepancy is likely to be explained by the different experi-mental approaches used in these studies. Waltz et al. (7) meas-ured binding of radiolabeled Vn to cell-surface uPAR. Wei et al.(30) measured binding of Vn to plastic-adsorbed uPAR. Høyer-Hansen et al. (9) studied the binding of Vn to antibody-immo-bilized suPAR by surface plasmon resonance. The commondenominator of these experimental approaches is that the mo-bility of uPAR is restricted by being bound to the cell surface,adsorbed to plastic, or immobilized on antibodies. Under theseconditions uPAR dimerization may occur but will be less dy-namic (bound to the cell surface) or even completely static(after adsorption to plastic or antibodies). In our binding assayuPAR is free to self-associate and to bind uPA and Vn.

The three domains of uPAR (D1, D2, and D3) belong to theLy-6/uPAR domain family that includes glycosylphosphatidyl-inositol-anchored single-domain membrane proteins (e.g.CD59, E48, and Ly-6) as well as a large number of secretedsingle domain �-neurotoxins (31). Although none of the mam-malian members of this domain family have been shown tooligomerize, the �-bungarotoxin subfamily of the �-neurotoxinsdimerize in solution through a strong homophilic interdomaininteraction (32). It is therefore tempting to speculate that oneor more of the uPAR domains may engage in intermolecularhomophilic contacts. Indeed, it has been proposed that thebinding of uPA to uPAR involves a binding pocket formed by atight intramolecular contact between domains 1 and 3 of uPAR(33). If true, this may leave domain 2 free to form intermolec-ular homophilic contacts. This may also suggest why Vn bind-ing requires pro-uPA binding.

uPA is composed of an amino-terminal GFD, a kringle do-main (KD), and carboxyl-terminal catalytic domain. Whereasthe ATF and GFD fragments of uPA have identical affinities foruPAR, only ATF efficiently promotes Vn binding (Fig. 4). Thissuggest that the uPA KD, which is present in ATF but absentin GFD, is required to induce uPAR dimerization. In accord-ance with this notion it was recently found that both the GFDand the KD of uPA are required for efficient induction ofchemotaxis in smooth muscle cells (34). However, the moleculebinding the KD still has to be determined. In fact, the exact roleof the KD in uPAR dimerization remains to be ascertained.

Several lines of evidence suggest that uPA promotes uPARbinding to Vn by changing the conformation of the receptorwithout itself directly participating in the interaction with Vn.First, overexpression of uPAR in cells expressing no uPA may

result in strong binding/adhesion to Vn (10, 30). Second, Vnbinds to immobilized uPAR in the absence of uPA (9, 30). Theevident correlation between Vn binding activity and uPARoligomerization presented in this study strongly suggests thatthe critical event that determines high affinity uPAR bindingto Vn is receptor dimerization and not simply ligand binding ora general conformational change.

REFERENCES

1. Johnsen, M., Lund, L. R., Romer, J., Almholt, K., and Dano, K. (1998) Curr.Opin. Cell Biol. 10, 667–671

2. Ploug, M., Rønne, E., Behrendt, N., Jensen, A. L., Blasi, F., and Danø, K.(1991) J. Biol. Chem. 266, 1926–1933

3. Ploug, M., Kjalke, M., Rønne, E., Weidle, U., Høyer-Hansen, G., and Danø, K.(1993) J. Biol. Chem. 268, 17539–17546

4. Ossowski, L., and Aguirre-Ghiso, J. A. (2000) Curr. Opin. Cell Biol. 12,613–620

5. Chapman, H. A. (1997) Curr. Opin. Cell Biol. 9, 714–7246. Blasi, F. (1997) Immunol. Today 18, 415–4177. Waltz, D. A., and Chapman, H. A. (1994) J. Biol. Chem. 269, 14746–147508. Higazi, A. A., Upson, R. H., Cohen, R. L., Manuppello, J., Bognacki, J., Henkin,

J., McCrae, K. R., Kounnas, M. Z., Strickland, D. K., Preissner, K. T.,Lawler, and Cines, D. B. (1996) Blood 88, 542–551

9. Høyer-Hansen, G., Behrendt, N., Ploug, M., Danø, K., and Preissner, K. T.(1997) FEBS Lett. 420, 79–85

10. Sidenius, N., and Blasi, F. (2000) FEBS Lett. 470, 40–4611. Behrendt, N., Jensen, O. N., Engelholm, L. H., Mortz, E., Mann, M., and Dano,

K. (2000) J. Biol. Chem. 275, 1993–200212. Boyle, M. D., Chiodo, V. A., Lawman, M. J., Gee, A. P., and Young, M. (1987)

J. Immunol. 139, 169–17413. Gudewicz, P. W., and Gilboa, N. (1987) Biochem. Biophys. Res. Commun. 147,

1176–118114. Del Rosso, M., Anichini, E., Pedersen, N., Blasi, F., Fibbi, G., Pucci, M., and

Ruggiero, M. (1993) Biochem. Biophys. Res. Commun. 190, 347–35215. Gyetko, M. R., Todd, R. F. R., Wilkinson, C. C., and Sitrin, R. G. (1994) J. Clin.

Invest. 93, 1380–138716. Resnati, M., Guttinger, M., Valcamonica, S., Sidenius, N., Blasi, F., and

Fazioli, F. (1996) EMBO J. 15, 1572–158217. Fazioli, F., Resnati, M., Sidenius, N., Higashimoto, Y., Appella, E., and Blasi,

F. (1997) EMBO J. 16, 7279–728618. Rao, N. K., Shi, G. P., and Chapman, H. A. (1995) J. Clin. Invest. 96, 465–47419. Kyhse Andersen, J. (1984) J. Biochem. Biophys. Methods 10, 203–20920. Ploug, M., Ellis, V., and Danø, K. (1994) Biochemistry 33, 8991–899721. Behrendt, N., Rønne, E., and Danø, K. (1996) J. Biol. Chem. 271, 22885–2289422. Ploug, M., Rahbek Nielsen, H., Nielsen, P. F., Roepstorff, P., and Danø, K.

(1998) J. Biol. Chem. 273, 13933–1394323. Ploug, M., Østergaard, S., Hansen, L. B., Holm, A., and Danø, K. (1998)

Biochemistry 37, 3612–362224. Bdeir, K., Kuo, A., Mazar, A., Sachais, B. S., Xiao, W., Gawlak, S., Harris, S.,

Higazi, A. A., and Cines, D. B. (2000) J. Biol. Chem. 275, 28532–2853825. Moser, T. L., Enghild, J. J., Pizzo, S. V., and Stack, M. S. (1995) Biochem. J.

307, 867–87326. Stoppelli, M. P., Corti, A., Soffientini, A., Cassani, G., Blasi, F., and Associan,

R. K. (1985) Proc. Natl. Acad. Sci. U. S. A. 82, 4939–494327. Appella, E., Robinson, E. A., Ullrich, S. J., Stoppelli, M. P., Corti, A., Cassani,

G., and Blasi, F. (1987) J. Biol. Chem. 262, 4437–444028. Shliom, O., Huang, M., Sachais, B., Kuo, A., Weisel, J. W., Nagaswami, C.,

Nassar, T., Bdeir, K., Hiss, E., Gawlak, S., Harris, S., Mazar, A., and Higazi,A. A. (2000) J. Biol. Chem. 275, 24304–24312

29. Liang, O. D., Chavakis, T., Kanse, S. M., and Preissner, K. T. (2001) J. Biol.Chem. 276, 28946–28953

30. Wei, Y., Waltz, D. A., Rao, N., Drummond, R. J., Rosenberg, S., and Chapman,H. A. (1994) J. Biol. Chem. 269, 32380–32388

31. Ploug, M., and Ellis, V. (1994) FEBS Lett. 349, 163–16832. Chiappinelli, V. A., and Lee, J. C. (1985) J. Biol. Chem. 260, 6182–618633. Ploug, M. (1998) Biochemistry 37, 16494–1650534. Mukhina, S., Stepanova, V., Traktouev, D., Poliakov, A., Beabealashvilly, R.,

Gursky, Y., Minashkin, M., Shevelev, A., and Tkachuk, V. (2000) J. Biol.Chem. 275, 16450–16458

Ligand-induced Urokinase Receptor Oligomerization27990

by guest on February 17, 2016http://w

ww

.jbc.org/D

ownloaded from

Nicolai Sidenius, Annapaola Andolfo, Riccardo Fesce and Francesco BlasiOligomerization

Urokinase Regulates Vitronectin Binding by Controlling Urokinase Receptor

2002, 277:27982-27990.J. Biol. Chem. 

  http://www.jbc.org/content/277/31/27982Access the most updated version of this article at

 Alerts:

  When a correction for this article is posted• 

When this article is cited• 

to choose from all of JBC's e-mail alertsClick here

  http://www.jbc.org/content/277/31/27982.full.html#ref-list-1

This article cites 34 references, 18 of which can be accessed free at

by guest on February 17, 2016http://w

ww

.jbc.org/D

ownloaded from


Recommended