14.01.2013 Views

Molecular and Cellular Biology of Plasminogen Activation

Molecular and Cellular Biology of Plasminogen Activation

Molecular and Cellular Biology of Plasminogen Activation

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

X I t h I n t e r n a t I o n a l w o r k S h o p o n<br />

<strong>Molecular</strong> <strong>and</strong> <strong>Cellular</strong> <strong>Biology</strong><br />

<strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong><br />

International Scientific<br />

Committee<br />

T. Antalis, Rockville, USA<br />

N. Behrendt, Copenhagen, Denmark<br />

F. Blasi, Milan, Italy<br />

P. Bock, Nashville, USA<br />

N. Brünner, Copenhagen, Denmark<br />

T. Bugge, Bethesda, USA<br />

P. DeClerk, Leuven, Belgium<br />

V. Ellis, Norwich, United Kingdom<br />

M. Gyetko, Ann Arbor, USA<br />

K. Hajjar, New York, USA<br />

M. Huang, Fujian, China<br />

E.K.O. Kruith<strong>of</strong>, Geneva, Switzerl<strong>and</strong><br />

K. Liu, Umeå, Sweden<br />

D. Loskut<strong>of</strong>f, La Jolla, USA<br />

V. Magdolen, München, Germany<br />

R. Medcalf, Box Hill, Australia<br />

D. Monard, Basel, Switzerl<strong>and</strong><br />

Y. Nagamine, Basel, Switzerl<strong>and</strong><br />

L. Ossowski, New York, USA<br />

L. Pathy, Budapest, Hungary<br />

16–20 June 2007<br />

C. Peterson, Knoxville, USA<br />

M. Ploug, Copenhagen, Denmark<br />

M. Ranson, Wollongong, Australia<br />

J. Rømer, Copenhagen, Denmark<br />

M.P. Stoppelli, Naples, Italy<br />

D. Strickl<strong>and</strong>, Rockville, USA<br />

A. Vaheri, Helsinki, Finl<strong>and</strong><br />

M. Wilczynska, Umeå, Sweden<br />

Organizing Committee<br />

P. Andreasen, Aarhus, Denmark<br />

T. Ny, Umeå, Sweden<br />

Workshop Coordinator<br />

E. Wolff, Urbana, USA<br />

Participants<br />

ConferenCe Centre Vår Gård<br />

SaltSJöbaden,Sweden<br />

130 participants from Australia, Austria,<br />

Belgium, Canada, China, Denmark,<br />

Germany, Italy, Japan, Mexico, Norway,<br />

Pol<strong>and</strong>, Russia, Spain, Sweden,<br />

Switzerl<strong>and</strong>, United Kingdom, <strong>and</strong> USA.


ii X I t h I n t e r n a t i o n a l W o r k s h o p o n


The XIth International Workshop on <strong>Molecular</strong><br />

<strong>and</strong> <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong><br />

is sponsored by<br />

Nobel Committee for Chemistry<br />

Swedish Cancer Society<br />

Swedish Research Council<br />

American Diagnostica Inc.<br />

The organizing committee would like to thank the organizers <strong>of</strong> the<br />

previous International Workshops <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> for<br />

generously contributing unspent funds.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> iii


Housing Grant Recipients<br />

Nina Ahlskog, Umeå, Sweden<br />

Daniela Alfano, Naples, Italy<br />

Lisbeth Andersen, Aarhus, Denmark<br />

Annapaola Andolfo, Milan, Italy<br />

Esther Ardite, Barcelona, Spain<br />

Rashna Balsara, Notre Dame, USA<br />

Nathalie Beaufort, Munich, Germany<br />

Julie Bødker, Aarhus, Denmark<br />

Patrick Brunner, Vienna, Austria<br />

Katharaina Bruno, Chicago, USA<br />

Blake Cochran, Wollongong, Australia<br />

David Croucher, Sydney, Australia<br />

Angels Diaz-Ramos, Barcelona, Spain<br />

Daniel Dupont, Aarhus, Denmark<br />

Monika Ehnman, Stockholm, Sweden<br />

Monika Ehart, Vienna, Austria<br />

Paola Franco, Naples, Italy<br />

Yongzhi Guo, Umeå, Sweden<br />

Peter Hägglöf, Cambridge, United Kingdom<br />

Jakob Harslund, Frederiksberg, Denmark<br />

Emir Henic, Lund, Sweden<br />

Karin Hultman, Gothenburg, Sweden<br />

Benedikte Jacobsen, Copenhagen, Denmark<br />

Lotte Jensen, Frederiksberg, Denmark<br />

Anna Juncker-Jensen, Copenhagen, Denmark<br />

Jodi Lee, Wollongong, Australia<br />

Shih-Hon Li, Urbana, USA<br />

Anna Lillis, Baltimore, USA<br />

Sergei Lobov, Wollongong, Australia<br />

Immacolata Longanesi Cattani, Naples, Italy<br />

Ida Katrine Lund, Copenhagen, Denmark<br />

Chris Madsen, Milan, Italy<br />

Rajani Maiya, New York, USA<br />

Lester Meissenheimer, Leuven, Belgium<br />

Judit Mihaly, Vienna, Austria<br />

Evelyn Nieves-Li, Urbana, USA<br />

Bjorn Olausson, Umeå, Sweden<br />

Mohan Pabba, Umeå, Sweden<br />

Justin Paul, New York, USA<br />

Valentina Pirazzoli, Milan, Italy<br />

Boris Pliyev, Moscow, Russia<br />

Gerald Prager, Vienna, Austria<br />

Patrycja Przygodzka, Umeå, Sweden<br />

Tomasz Przygodzki, Umeå, Sweden<br />

Aless<strong>and</strong>ro Salvi, Brescia, Italy<br />

Morten Rasch, Copenhagen, Denmark<br />

Birgitte Rønø, Copenhagen, Denmark<br />

Gian Maria Sarra Ferraris, Milan, Italy<br />

Rima Sulniute, Umeå, Sweden<br />

Berta Vidal, Barcelona, Spain<br />

Patrik Wahlberg, Umeå, Sweden<br />

Ying Wei, San Francisco, USA<br />

Malgorzata Wygrecka, Giessen, Germany<br />

Wendy Xolalpa, Mexico, Mexico<br />

Aiwu Zhou, New York, USA<br />

iv X I t h I n t e r n a t i o n a l W o r k s h o p o n


Table <strong>of</strong> Contents<br />

Saturday, 16 June 2007 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2<br />

Sunday, 17 June 2007 ..............................................................2<br />

Session 1—Vitronectin <strong>and</strong> the <strong>Plasminogen</strong> <strong>Activation</strong> System ............................2<br />

Session 2—Proteases, Inhibitors, Receptors, <strong>and</strong> Substrates in Vascular <strong>Biology</strong> . . . . . . . . . . . . . .3<br />

Session 3—Cell Signalling Upstream <strong>and</strong> Downstream <strong>of</strong> the<br />

<strong>Plasminogen</strong> <strong>Activation</strong> System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3<br />

Monday, 18 June 2007 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4<br />

Session 4—Proteases, Inhibitors, Receptors, <strong>and</strong> Substrates in Neurobiology<br />

<strong>and</strong> Neuronal Pathology ..............................................................4<br />

Session 5—Proteases, Inhibitors, Receptors, <strong>and</strong> Substrates in Normal Development ..........5<br />

Session 6—The <strong>Plasminogen</strong> <strong>Activation</strong> System <strong>and</strong> Programmed Cell Death ................5<br />

Tuesday, 19 June 2007 .............................................................6<br />

Session 7—Proteases, Inhibitors, Receptors, <strong>and</strong><br />

Substrates in Tissue Repair <strong>and</strong> Tissue Remodelling ......................................6<br />

Session 8—Proteases, Inhibitors, Receptors, <strong>and</strong><br />

Substrates in Inflammation <strong>and</strong> Infectious Diseases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6<br />

Wednesday, 20 June 2007 ..........................................................7<br />

Session 9—Proteases, Inhibitors, Receptors, <strong>and</strong> Substrates in Cancer .......................7<br />

Session 10—Proteases, Inhibitors, <strong>and</strong> Receptors as Therapeutic Targets <strong>and</strong><br />

Therapeutic Proteins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7<br />

Poster Presentations ................................................................9<br />

Abstracts for Oral Presentations ...................................................13<br />

Abstracts for Poster Presentations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .63<br />

Author Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .129<br />

Attendee List .....................................................................132<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 1


Saturday, 16 June 2007<br />

14:00–18:00 Registration<br />

17:30–17:40 Welcome <strong>and</strong> Meeting Overview<br />

Peter Andreasen <strong>and</strong> Tor Ny<br />

17:40–18:25 Popular lecture: “Carl Linnaeus—The man who gave all <strong>of</strong> us our names”<br />

Anna Rask-Andersen<br />

18:30–20:00 Dinner<br />

20:15–21:00 Lecture: “Control <strong>of</strong> invasive growth by plasminogen-related growth factors<br />

(HGF <strong>and</strong> MSP)”<br />

Paolo Michieli <strong>and</strong> Paolo Comoglio<br />

Sunday, 17 June 2007<br />

7:00–8:30 Breakfast<br />

X I t h I n t e r n a t I o n a l w o r k S h o p o n<br />

<strong>Molecular</strong> <strong>and</strong> <strong>Cellular</strong> <strong>Biology</strong><br />

<strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong><br />

8:30–10:15 Session 1—Vitronectin <strong>and</strong> the <strong>Plasminogen</strong> <strong>Activation</strong> System<br />

Session Chairs: Michael Ploug <strong>and</strong> Mingdong Huang<br />

8:30–8:35 Introduction by Session Chairs<br />

8:35–8:55 001 • Defining the Native Disulfide Topology in the SMB Domain <strong>of</strong> Human<br />

Vitronectin • Li X, Zou G, Yuan W, Lu W*<br />

8:55–9:15 002 • The Vitronectin Binding Site on the Urokinase Receptor Comprises Residues<br />

from Both Domain I <strong>and</strong> the Flanking Interdomain Linker Region • Gårdsvoll H <strong>and</strong><br />

Ploug M*<br />

9:15–9:35 003 • uPAR-induced Cell Adhesion <strong>and</strong> Migration: Vitronectin Provides the Key •<br />

Madsen CD*, Sarra Ferraris GM, Andolfo A, Cunningham O, Sidenius N<br />

9:35–9:55 004 • Interactions between PAI-1 <strong>and</strong> Vitronectin: Two Proteins, Two Sites, <strong>and</strong> Two<br />

Phases • Schar CR, Jensen JK, Blouse GE, Minor KH, Andreasen PA, Peterson CB*<br />

9:55–10:15 005 • How Does Vitronectin Accelerate PAI-1’s Protease Inhibition? • Zhou A* <strong>and</strong><br />

Wei Z<br />

Relevant Posters—051–056<br />

2 X I t h I n t e r n a t i o n a l W o r k s h o p o n


10:15–10:35 Break<br />

10:35–12:20 Session 2—Proteases, Inhibitors, Receptors, <strong>and</strong> Substrates in Vascular <strong>Biology</strong><br />

Session Chairs: Marie Ranson <strong>and</strong> Thomas Bugge<br />

10:35–10:40 Introduction by Session Chairs<br />

10:40–11:00 006 • The Interaction between Tissue-type <strong>Plasminogen</strong> Activator <strong>and</strong> the Low<br />

Density Lipoprotein Receptor-Related Protein Induces <strong>Activation</strong> <strong>of</strong> the NF-kB<br />

Pathway during Cerebral Ischemia • Yepes M*, Brotzge XH, Polavarapu R<br />

11:00–11:20 007 • Unique Secretory Dynamics <strong>of</strong> Tissue <strong>Plasminogen</strong> Activator (tPA) Is<br />

Beneficial to Maintain Fibrinolytic Activity on Cell Surface • Suzuki Y*, Ihara H,<br />

Mogami H, Urano T<br />

11:20–11:40 008 • The Macrophage Low-Density Lipoprotein Receptor Related Protein (LRP)<br />

Modulates Murine Lipoprotein Metabolism • Lillis AP*, Mikhailenko I, Robinson S,<br />

Migliorini M, Battey F, Pizzo SV, Strickl<strong>and</strong> DK<br />

11:40–12:00 009 • Annexin 2 Mediates <strong>Plasminogen</strong>-Dependent Recruitment <strong>of</strong> Neovascular<br />

Mural Cells in Lymphoma Angiogenesis • Ling Q, Ruan J, Yan L, Sui G-Z, Deora AB,<br />

Church S, Cohen-Gould L, Rafii S, Lyden D, Hajjar KA*<br />

12:00–12:20 010 • <strong>Activation</strong> <strong>of</strong> Latent PDGF-CC by Tissue <strong>Plasminogen</strong> Activator Impairs<br />

Blood Brain Barrier Integrity during Ischemic Stroke • Su EJ, Fredriksson L,<br />

Geyer M, Folestad E, Cale J, Mann K, Gao Y, Pietras K, Andreé J, Yepes M, Strickl<strong>and</strong> DK,<br />

Betsholtz C, Eriksson U, Lawrence DA*<br />

12:20–12:30 Group photo<br />

Relevant Posters—057–060<br />

12:30–15:00 Lunch <strong>and</strong> free time<br />

15:00–17:30 Poster Session A—Posters 051–084<br />

17:30–19:00 Dinner<br />

19:00–21:10 Session 3—Cell Signalling Upstream <strong>and</strong> Downstream <strong>of</strong> the <strong>Plasminogen</strong><br />

<strong>Activation</strong> System<br />

Session Chairs: Francesco Blasi <strong>and</strong> Egbert K.O. Kruith<strong>of</strong><br />

19:00–19:05 Introduction by Session Chairs<br />

19:05–19:25 011 • <strong>Plasminogen</strong> Activator Inhibitor-1 Gene Regulation: Cross Talk between<br />

Hypoxia <strong>and</strong> Insulin Signalling • Flugel D <strong>and</strong> Kietzmann T*<br />

19:25–19:45 012 • Role <strong>of</strong> Rho GTPases <strong>and</strong> p38 MAP Kinase in the Regulation <strong>of</strong> t-PA <strong>and</strong><br />

PAI-1 Expression in Cultured Human Endothelial Cells • Fish RJ*, Dunoyer-Geindre<br />

S, Kruith<strong>of</strong> EKO<br />

19:45–20:05 013 • Urokinase Receptor/a5b1 Integrin Interaction <strong>and</strong> Signaling in Cancer Cells •<br />

Wei Y*, Tang CH, Kim Y, Robillard L, Kugler MC, Hill M, Brumwell A, Chapman HA<br />

20:05–20:10 Break<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 3


20:10–20:30 014 • Activated Human Neutrophils Rapidly Release the Chemotactically Active<br />

D2D3 Form <strong>of</strong> the Urokinase-type <strong>Plasminogen</strong> Activator Receptor (uPAR/CD87) •<br />

Pliyev BK* <strong>and</strong> Tkachuk VA<br />

20:30–20:50 015 • PDGF-DD Bioavailability Is Regulated by the uPA/uPAR System:<br />

Implications for Tumor Growth • Ehnman M*, Li H, Fredriksson L, Eriksson U<br />

20:50–21:10 064 • Detection <strong>and</strong> Prevention <strong>of</strong> Hepatic Fibrosis Targeting Proteolytic TGF-b<br />

<strong>Activation</strong> Reaction • Kojima S<br />

Relevant Posters—061–076<br />

Monday, 18 June 2007<br />

7:00–8:30 Breakfast<br />

8:30–11:00 Session 4—Proteases, Inhibitors, Receptors, <strong>and</strong> Substrates in Neurobiology <strong>and</strong><br />

Neuronal Pathology<br />

Session Chairs: Denis Monard <strong>and</strong> Dan Lawrence<br />

8:30–8:35 Introduction by Session Chairs<br />

8:35–8:55 017 • A Novel Neuronal Death Pathway Triggered By Excess Tissue <strong>Plasminogen</strong><br />

Activator • Li J*, Snyder EY, Sidman RL<br />

8:55–9:15 018 • The Inhibitor <strong>of</strong> Serine Proteases Protease Nexin-1 <strong>and</strong> its Receptor LRP<br />

Modulate SHH Signalling during Cerebellar Development • Vaillant C, Michos O,<br />

Orolicki S, Brellier F, Taieb S, Moreno E, Té H, Zeller R, Monard D*<br />

9:15–9:35 019 • Fibrin Deposition Accelerates Neurovascular Damage <strong>and</strong><br />

Neuroinflammation in Mouse Models <strong>of</strong> Alzheimer’s Disease • Paul J* <strong>and</strong><br />

Strickl<strong>and</strong> S<br />

9:35–9:55 020 • Anxiety-like Behavior <strong>and</strong> Impaired Fear Extinction in Mice with Altered<br />

Control <strong>of</strong> Extracellular Brain Proteolytic Activity • Meins M*, Herry C, Moreno E,<br />

Fischer C, Lüthi A, Monard D<br />

9:55–10:00 Break<br />

10:00–10:20 021 • Tissue <strong>Plasminogen</strong> Activator Is Co-Packaged <strong>and</strong> Co-Transported to<br />

Synaptic Sites with a Key Neuromodulator Associated with Synaptic Plasticity •<br />

Lochner JE*, Spangler E, Schuttner LC, Scalettar BA<br />

10:20–10:40 022 • Tissue <strong>Plasminogen</strong> Activator Modulates <strong>Cellular</strong> <strong>and</strong> Behavioral Response to<br />

Cocaine • Maiya R*, Zhou Y, Norris EH, Kreek MJ, Strickl<strong>and</strong> S<br />

10:40–11:00 023 • Characterisation <strong>of</strong> the Pathway <strong>of</strong> Polymerisation <strong>of</strong> Wildtype Neuroserpin<br />

<strong>and</strong> the Ser49Pro Mutant that Underlies the Dementia FENIB • Hägglöf P*,<br />

Belorgey D, Karlsson-Li S, Sharp LK, Lomas DA<br />

Relevant Poster—077<br />

4 X I t h I n t e r n a t i o n a l W o r k s h o p o n


11:00–11:20 Break<br />

11:20–12:25 Session 5—Proteases, Inhibitors, Receptors, <strong>and</strong> Substrates in Normal<br />

Development<br />

Session Chairs: Kui Liu <strong>and</strong> Leif Lund<br />

11:20–11:25 Introduction by Session Chairs<br />

11:25–11:45 024 • Matriptase Is an Essential Inhibitory Target for Hepatocyte Growth Factor<br />

Activator Inhibitor-1 during both Embryonic Development <strong>and</strong> Postnatal Life •<br />

Szabo R*, Molinolo A, List K, Bugge TH<br />

11:45–12:05 025 • Mice with very low Matriptase Are Viable <strong>and</strong> Phenocopy Human Autosomal<br />

Ichthyosis with Hypotrichosis Syndrome • List K*, Currie B, Scharschmidt T, Szabo R,<br />

Molinolo A, Shireman J, Segre J, Bugge TH<br />

12:05–12:25 026 • Role <strong>of</strong> Urokinase-Receptor in Hematopoietic Stem Cell Trafficking •<br />

Montuori N, Selleri C, Ricci P, Visconte V,Carriero MV, Rotoli B, Rossi G, Ragno P*<br />

Relevant Posters—078–084<br />

12:25–17:30 Lunch <strong>and</strong> Free Time<br />

17:30–18:30 Dinner<br />

18:30–20:15 Session 6—The <strong>Plasminogen</strong> <strong>Activation</strong> System <strong>and</strong> Programmed Cell Death<br />

Session Chairs: Patrizia Stoppelli <strong>and</strong> Bernd Binder<br />

18:30–18:35 Introduction by Session Chairs<br />

18:35–18:55 027 • <strong>Plasminogen</strong> Activator Inhibitor-1, PAI-1, Regulates the Akt Survival Pathway<br />

• Rømer MU, Larsen L, Offenberg H, Brünner N, Lademann U*<br />

18:55–19:15 028 • A Host <strong>Plasminogen</strong> Activator Inhibitor-1 Deficiency Promotes Proliferation<br />

<strong>and</strong> Resistance to Apoptosis by <strong>Activation</strong> <strong>of</strong> the PI3-K/Akt Pathway in Endothelial<br />

Cells • Balsara RD*, Castellino FJ, Ploplis VA<br />

19:15–19:35 029 • Bomapin Is a Redox-Regulated Serpin which Stabilizes Retinoblastoma<br />

Protein during Apoptosis <strong>and</strong> Increases Proliferation <strong>of</strong> Leukemia Cells •<br />

Przygodzka P, Olausson B, Tengel Y, Larsson G, Wilczynska M*<br />

19:35–19:55 030 • Hepsin as a Cell Survival Factor • Qiu D, Owen K, Edwards DR, Ellis V*<br />

19:55–20:15 031 • Urokinase (uPA) Protects Endothelial Cell against Apoptosis by Upregulating<br />

the X-Linked Inhibitor <strong>of</strong> Apoptosis Protein (XIAP) •Prager GW*, Koschelnick Y,<br />

Mihaly J, Brunner P, Binder BR<br />

Relevant Posters—085–087<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 5


Tuesday, 19 June 2007<br />

7:00–8:30 Breakfast<br />

8:30–9:55 Session 7—Proteases, Inhibitors, Receptors, <strong>and</strong> Substrates in Tissue Repair <strong>and</strong><br />

Tissue Remodelling<br />

Session Chairs: Katherine Hajjar <strong>and</strong> Dudley Strickl<strong>and</strong><br />

8:30–8:35 Introduction by Session Chairs<br />

8:35–8:55 032 • Pro-fibrinolytic Effects <strong>of</strong> Metalloproteinases during Skin Wound Healing in<br />

the Absence <strong>of</strong> <strong>Plasminogen</strong> • Lund LR*, Green KA, Almholt K, Ploug M, Bugge TH,<br />

Rømer J<br />

8:55–9:15 033 • Complementary Roles <strong>of</strong> Intracellular <strong>and</strong> Pericellular Collagen Degradation<br />

Pathways in Mesenchymal Cell Survival <strong>and</strong> Proliferation • Wagenaar-Miller RA,<br />

Engelholm LH, Gavard J, Yamada S, Gutkind JS, Behrendt N, Holmbeck K, Bugge TH*<br />

9:15–9:35 034 • Interplay between MMPs <strong>and</strong> the Endocytic Collagen Receptor, uPARAP/<br />

Endo180, in Collagen Degradation • Behrendt N*, Madsen DH, Ingvarsen S, Hillig T,<br />

Wagenaar-Miller R, Kjøller L, Gårdsvoll H, Høyer-Hansen G, Bugge TH, Engelholm LH<br />

9:35–9:55 035 • The Urokinase Receptor Ko Mice Have Reduced Keratinocytes Proliferation<br />

<strong>and</strong> Migration during Wound Healing <strong>and</strong> Are Protected in a Skin Carcinogenesis<br />

Protocol • D’Alessio S, Mazzieri, R, Gerasi L, Blasi F*<br />

9:55–10:15 Break<br />

Relevant Poster—088<br />

10:15–12:00 Session 8—Proteases, Inhibitors, Receptors, <strong>and</strong> Substrates in Inflammation <strong>and</strong><br />

Infectious Diseases<br />

Session Chairs: Toni Antalis <strong>and</strong> Vincent Ellis<br />

10:15–10:20 Introduction by Session Chairs<br />

10:20–10:40 036 • Distinct Roles <strong>of</strong> Plasmin in Staphylococcus aureus-induced Sepsis <strong>and</strong><br />

Infection Models • Guo Y*, Li J, Hagström E, Ny T<br />

10:40–11:00 037 • Proteolytic <strong>Activation</strong> <strong>of</strong> the Human Urokinase/Plasmin System by<br />

Staphylococcus aureus • Beaufort N*, Wojciechowski P, Sommerh<strong>of</strong>f CP, Schmitt M,<br />

Potempa J, Magdolen V<br />

11:00–11:20 038 • The Maintenance <strong>of</strong> High Affinity <strong>Plasminogen</strong> Binding by PAM Variants<br />

from Group A Streptococci Is Mediated by Conserved Arg <strong>and</strong> His Residues in Both<br />

the A1 <strong>and</strong> A2 Repeat Domains • Ranson M*, S<strong>and</strong>erson-Smith ML, Walker MJ, Fu Q,<br />

Castellino FJ, Prorok M<br />

11:20–11:40 039 • <strong>Plasminogen</strong> Activator Inhibitor-1 (PAI-1) Is an Inhibitor <strong>of</strong> Factor VII-<br />

Activating Protease in Patients with Acute Respiratory Distress Syndrome •<br />

Wygrecka M*, Morty RE, Markart P, Kanse SM, Andreasen PA, Wind T, Guenther A,<br />

Preissner KT<br />

6 X I t h I n t e r n a t i o n a l W o r k s h o p o n


11:40–12:00 040 • Improved Muscle Regeneration in PAI-1-deficient Mice Is Associated with an<br />

Enhanced Inflammatory Response <strong>and</strong> Reduced Fibrin Deposition after Injury •<br />

Ardite E*, Vidal B, Jardí M, González B, Muñoz-Cánoves P<br />

Relevant Posters—089–095<br />

12:00–12:20 Presentation <strong>of</strong> the 2009 <strong>Plasminogen</strong> <strong>Activation</strong> Meeting<br />

12:20–14:30 Lunch <strong>and</strong> Free Time<br />

14:30–17:00 Poster Session B—Posters 085–116<br />

17:00 Reception <strong>and</strong> Gala Banquet <strong>and</strong> Boat Trip<br />

Wednesday, 20 June 2007<br />

7:00–8:30 Breakfast<br />

8:30–9:55 Session 9—Proteases, Inhibitors, Receptors, <strong>and</strong> Substrates in Cancer<br />

Session Chairs: Keld Danø <strong>and</strong> Thomas Kietzmann<br />

8:30–8:35 Introduction by Session Chairs<br />

8:35–8:55 041 • Invasion <strong>and</strong> Metastasis <strong>of</strong> Carcinoma Cells Is Prevented by Urokinase-<br />

Derived Antagonists <strong>of</strong> avb5 Integrin <strong>Activation</strong> • Franco P, Vocca I, Alfano D,<br />

Votta G, Carriero MV, Estrada Y, Netti PA, Ossowski L, Stoppelli MP*<br />

8:55–9:15 042 • Urokinase Receptor/Integrin Interactions in Lung Tumor Development •<br />

Tang CH, Hill M, Kim Y, Wei Y, Chapman HA*<br />

9:15–9:35 043 • uPA <strong>and</strong> uPAR Expressing Stromal Cells Accompany the Transition to<br />

Invasive Breast Cancer • Nielsen BS*, Rank F, Illemann M, Lund LR, Danø K<br />

9:35–9:55 044 • Generation <strong>of</strong> the Malignant Phenotype in HT-1080 Tumor Cells by<br />

PAI-1 Involves Modulation <strong>of</strong> Proteasomal Activity <strong>and</strong> Phosphatases • Mihaly J*,<br />

Carroll VA, Breuss JM, Prager GW, Binder BR<br />

9:55–10:15 Break<br />

Relevant Posters—096–106<br />

10:15–12: 20 Session 10—Proteases, Inhibitors, <strong>and</strong> Receptors as Therapeutic Targets <strong>and</strong><br />

Therapeutic Proteins<br />

Session Chairs: Gunilla Høyer-Hansen <strong>and</strong> Ann Gils<br />

10:15–10:20 Introduction by Session Chairs<br />

10:20–10:40 045 • PEGylated DX-1000: Pharmacokinetics, Anti-Tumor <strong>and</strong> Anti-Metastatic<br />

Effects <strong>of</strong> a Specific Plasmin Inhibitor • Devy L*, Rabbani SA, Stochl M, Ruskowski M,<br />

Mackie I, Naa L, Toews M, van Gool R, Chen J, Ley A, Ladner RC, Dransfield DT,<br />

Henderikx P<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 7


10:40–11:00 046 • Cytotoxic Potential <strong>of</strong> a Novel uPA-activity Dependent <strong>and</strong> EGF Receptor<br />

Targeting Pro-drug • Rønø B*, Kim GB, Liu S, Kristjansen PEG, Neville DM, Leppla SH,<br />

Bugge TH, Rømer J<br />

11:00–11:20 047 • Inhibition <strong>of</strong> Mouse uPA Activity by Mouse Monoclonal Antibodies in vitro<br />

<strong>and</strong> in vivo • Lund IK*, Jögi A, Behrendt N, Ploug M, Gårdsvoll H, Lund LR, Rømer J,<br />

Høyer-Hansen G<br />

11:20–11:40 048 • Crystal Structure <strong>of</strong> Human Urokinase Complexed with a Cyclic Peptidyl<br />

Inhibitor, uPAin-1 • Zhao G, Yuan C, Bian C, Wind T, Andreasen PA, Huang M*<br />

11:40–12:00 049 • Discovery <strong>of</strong> a Novel Zymogen Targeting Inhibitor <strong>of</strong> Urokinase-type<br />

<strong>Plasminogen</strong> Activator: Evidence for Structural Flexibility <strong>of</strong> the Protease Domain •<br />

Blouse GE, Bøtkjær KA, Deryugina EI, Kjelgaard S, Byszuk O, Mortensen KK, Quigley JP,<br />

Andreasen PA*<br />

12:00–12:20 050 • A Novel Type <strong>of</strong> Agent Blocking the Association <strong>of</strong> uPA to its Receptor<br />

uPAR: uPA-binding Aptamers • Dupont DM*, Madsen JB, Kjems J, Andreasen PA<br />

Relevant Posters—107–116<br />

12:20 Lunch <strong>and</strong> departure<br />

8 X I t h I n t e r n a t i o n a l W o r k s h o p o n


Poster Presentations<br />

• Poster numbers 51–84 will be<br />

displayed from Saturday evening<br />

through Monday morning.<br />

• Presenters <strong>of</strong> poster numbers 51–84<br />

will present in Poster Session A on<br />

Sunday, 17 June from 15:00–17:30.<br />

• Poster numbers 85–116 will be<br />

displayed from Monday afternoon<br />

through Wednesday morning.<br />

• Presenters <strong>of</strong> poster numbers 85–116<br />

will present in Poster Session B on<br />

Tuesday, 19 June from 14:30–17:00.<br />

• Presenters should be present at their<br />

posters during the first 2 hours <strong>of</strong> the<br />

poster session.<br />

051 • Identification <strong>and</strong> Analysis <strong>of</strong> the Vn<br />

Binding Site in Mouse uPAR • Pirazzoli V*,<br />

Andolfo AP, Madsen CD, Sidenius N<br />

052 • Novel uPAR Binding Site in Vitronectin<br />

• Andolfo A* <strong>and</strong> Sidenius N<br />

053 • Dissecting Serpin-protease Reaction<br />

Pathways by the Use <strong>of</strong> Monoclonal<br />

Antibodies • Bødker JS*, Blouse GE,<br />

Dupont DM, Andreasen PA<br />

054 • PAI-1-vitronectin Interactions Involve<br />

an Extended Binding Surface <strong>and</strong> Mutual<br />

Conformational Rearrangements • Blouse GE,<br />

Peterson CB, Dupont DM, Ploug M, Gårdsvoll H,<br />

Schar CR, Perron MJ, Minor KH, Shore JD,<br />

Andreasen PA*<br />

055 • Intact (non-cleavable) Cell-surface<br />

u-PAR Accelerates Clearance <strong>of</strong><br />

tcu-PA:PAI-1:u-PAR Complexes <strong>and</strong><br />

Subsequent Re-surfacing <strong>of</strong> Intact <strong>and</strong><br />

Functional u-PAR • Nieves-Li EC* <strong>and</strong><br />

Manch<strong>and</strong>a N<br />

056 • The Central b-sheet <strong>of</strong> PAI-1<br />

Demonstrates Two Dynamically Distinct<br />

Regions • Li S*, Lawrence DA, Schwartz BS<br />

057 • Regulation <strong>of</strong> Cancer Cell Plasmin<br />

Generation by Annexin A2-S100A10<br />

Heterotetramer (AIIt) • Waisman DM*<br />

058 • <strong>Plasminogen</strong> Activator Inhibitor Type 2<br />

Binds to S100A10 in Annexin 2 Heterotetramer<br />

<strong>and</strong> Prevents Annexin 2-dependent Plasmin<br />

In-site Formation by Inhibiting tPA • Lobov S*,<br />

Croucher D, Ranson M<br />

059 • Underst<strong>and</strong>ing the Structural Basis<br />

<strong>of</strong> the Differential-Receptor-Mediated<br />

Endocytosis Mechanisms <strong>of</strong> PAI-1 <strong>and</strong> PAI-2<br />

in Cancer • Cochran BJ*, Lobov S, Croucher D,<br />

Ranson M<br />

060 • A Low-glycemic-index Diet Reduces<br />

Plasma PAI-1 Activity in Overweight Women<br />

• Jensen L*, Krog-Mikkelsen I, Sloth B, Flint A,<br />

Astrup A, Raben A, Tholstrup T, Brünner N<br />

061 • Urokinase Receptor-independent<br />

Signalling <strong>of</strong> the Urokinase-type <strong>Plasminogen</strong><br />

Activator via Phosporylation <strong>of</strong> STAT1 • Ehart<br />

M* <strong>and</strong> Binder BR<br />

062 • Urokinase Receptor Promotes Neo-<br />

Angiogenesis through its Ser88-Arg-Ser-Arg-<br />

Tyr92 Chemotactic Sequence • Longanesi-<br />

Cattani I*, Bifulco K, Cantelmo AR, Di Carluccio<br />

G, Spina R, Liguori E, Stoppelli MP, Carriero MV<br />

063 • The Density Enhanced Phosphatase 1<br />

(DEP-1) Down-Modulates Urokinase Receptor<br />

(uPAR) Surface Expression in Confluent<br />

Endothelial Cells • Brunner PM*, Heier PC,<br />

Prager GW, Mihaly J, Priglinger U, Binder BR<br />

064 • Detection <strong>and</strong> Prevention <strong>of</strong> Hepatic<br />

Fibrosis Targeting Proteolytic TGF-b<br />

<strong>Activation</strong> Reaction • Kojima S*<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 9


065 • Domain 1 <strong>of</strong> uPAR Is Required for its<br />

Morphological <strong>and</strong> Functional b2 Integrinmediated<br />

Connection with Actin Cytoskeleton<br />

in Human Endothelial Cells • Del Rosso M*,<br />

Fibbi G, Margheri F, Serratì S, Pucci M,<br />

Manetti M, Ibba-Manneschi L<br />

066 • The uPA/uPAR/Vn Pathway <strong>of</strong><br />

Signaling to MAPK-activation • Sarra<br />

Ferraris GM*, Madsen C, Sidenius N<br />

067 • Regulation <strong>of</strong> tPA <strong>and</strong> PAI-1 Gene<br />

Expression in Astrocytes • Hultman K*,<br />

Tjärnlund-Wolf A, Blomstr<strong>and</strong> F, Nilsson M,<br />

Medcalf R, Jern C<br />

068 • Identification <strong>of</strong> a Mitotic Epitope in the<br />

Domain 2 <strong>of</strong> the Urokinase Receptor (uPAR)<br />

• Degryse B*, Eden G, Arnaudova R, Furlan F,<br />

Blasi F<br />

069 • Vitronectin Inhibits <strong>Plasminogen</strong><br />

Activator Inhibitor-1 (PAI-1)-Induced<br />

Chemotaxis by Blocking PAI-1 Binding to<br />

the LDL Receptor-Related Protein • Neels JG,<br />

Kamikubo Y, Degryse B*<br />

070 • Estradiol Inhibits EGF-induced Cell<br />

Migration <strong>and</strong> uPAR Expression in Estrogen<br />

Receptor-a Negative, GPR30 Positive Ovarian<br />

Cancer Cells • Henic E*, Noskova V, Høyer-<br />

Hansen G, Hansson S, Casslén B<br />

071 • Methylation <strong>of</strong> the PAI-1 Gene in Oral<br />

Squamous Cell Carcinomas <strong>and</strong> Normal Oral<br />

Mucosa • Gao S, Krogdahl A, Sørensen JA,<br />

Dabelsteen E, Andreasen PA*<br />

072 • Urokinase Signaling through Its<br />

Receptor Promotes Invasiveness <strong>and</strong><br />

Metastasis <strong>of</strong> Pancreatic Cancer Cells •<br />

Xue A*, Xue M, Jackson C, Song E, Allen BJ,<br />

Smith RC<br />

073 • Tissue <strong>Plasminogen</strong> Activator Induces<br />

Cell Proliferation in Pancreatic Cancer by<br />

a Non-catalytic Mechanism that Requires<br />

ERK1/2 <strong>Activation</strong> through Epidermal<br />

Growth Factor Receptor <strong>and</strong> Annexin A2 •<br />

Ortiz-Zapater E, Peiró S, Roda O, Corominas JM ,<br />

Aguilar S, Ampurdanés C, Real FX, Navarro P*<br />

074 • Modulation <strong>of</strong> Lung Carcinoma Cell<br />

Lines <strong>and</strong> Primary Cultures Migration by<br />

uPA-Derived <strong>and</strong> EGFR Inhibitors • Franco P*,<br />

Mancini A, Votta G, Caputi M, Stoppelli MP<br />

075 • A Novel Role <strong>of</strong> Ku80 in Regulation<br />

<strong>of</strong> PAI-1 Gene Expression in Migrating<br />

Endothelial Cells Induced by Thymosin b4<br />

• Bednarek R*, Boncela J, Smolarczyk K,<br />

Cierniewski CS<br />

076 • Signaling Pathway Involved in<br />

Inhibition <strong>of</strong> PAI-1 Expression by CNP in<br />

Endothelial Cells • Jerczynska H*, Cierniewski<br />

CS, Pawlowska Z<br />

077 • Interaction <strong>of</strong> Alzheimer’s Amyloid<br />

b-peptide (Ab) 1-40 with PAI-2 • Pabba M*,<br />

Przygodzki T, Malisauskas M, Ol<strong>of</strong>sson A,<br />

Morozova-Roche L, Wilczynska M, Ny T<br />

078 • The Subcellular Itinerary <strong>of</strong> Hepatocyte<br />

Growth Factor Activator Inhibitor-1 in MDCK<br />

Cells • Godiksen S, Selzer-Plon J, Pedersen EDK,<br />

Borger Rasmussen H, Bugge TH, Vogel LK*<br />

079 • Evidence for a Matriptase-Prostasin<br />

(CAP1/PRSS8) Serine Protease Zymogen-<br />

Cascade-Regulating Epithelial Differentiation<br />

• List K*, Netzel-Arnett S, Currie B, Szabo R,<br />

Molinolo A, Antalis TM, Bugge TH<br />

080 • The <strong>Plasminogen</strong> <strong>Activation</strong> System<br />

in Monocytic Cell Differentiation <strong>and</strong><br />

Proliferation: Potential Target for <strong>Plasminogen</strong><br />

<strong>Activation</strong> Inhibitor Type- 2-Based<br />

Therapeutics • Lee JA*, Croucher DR, Ranson M<br />

081 • Identification <strong>and</strong> Localization <strong>of</strong><br />

Novel Serine Proteases in the Mouse Ovary •<br />

Wahlberg P*, Nyl<strong>and</strong>er Å, Kui L, Ny T<br />

082 • Alpha-enolase/<strong>Plasminogen</strong> Binding Is<br />

Required during Myogenesis in vitro <strong>and</strong> in<br />

vivo • Diaz-Ramos A*, Llorens A, Luque T,<br />

López-Alemany R<br />

083 • The Serpinb8 Is Alternatively Spliced<br />

to the Known Long Form <strong>and</strong> a Novel Short<br />

Form • Olausson B*, Przygodzka P, Dahl L,<br />

Carlsson L, Wilczynska M<br />

10 X I t h I n t e r n a t i o n a l W o r k s h o p o n


084 • Characterization <strong>of</strong> a Combined PAI-1<br />

<strong>and</strong> TIMP-1 Gene-deficient Mouse Model •<br />

Harslund J*, Nielsen OL, Brünner N, Offenberg H<br />

085 • Dual Role <strong>of</strong> the uPA/uPAR System in<br />

Apoptosis <strong>of</strong> Mesangial Cells <strong>and</strong> Diabetic<br />

Nephropathy • Tkachuk N, Tkachuk S*, Kiyan J,<br />

Shushakova N, Haller H, Dumler I<br />

086 • Phenotypic Consequences <strong>of</strong><br />

<strong>Plasminogen</strong> Activator Inhibitor-1 Gene<br />

Ablation on STAT1 <strong>Activation</strong> <strong>and</strong> Cell Cycle<br />

Progression in Proliferating Endothelial<br />

Cells • Balsara RD*, Morin SJ, Meyer CA,<br />

Castellino FJ, Ploplis VA<br />

087 • Urokinase <strong>and</strong> its Receptor as Novel<br />

C-Myc Target Genes Affecting Cell Migration<br />

<strong>and</strong> Apoptosis • Alfano D*, Iaccarino I,<br />

Stoppelli MP<br />

088 • Effect <strong>of</strong> <strong>Plasminogen</strong> on Cell Migration<br />

Using an in vitro Wound Model • Sulniute R*,<br />

Li J, Ny T<br />

089 • uPA, but not its Receptor uPAR, Is<br />

Necessary for Experimentally-induced <strong>and</strong><br />

Pathological Muscle Regeneration • Vidal B*,<br />

Serrano AL, Jardí M, Suelves M, Muñoz-<br />

Cánoves P<br />

090 • Urokinase-type <strong>Plasminogen</strong> Activator<br />

Deficiency Strongly Attenuates Ischemia<br />

Reperfusion Injury <strong>and</strong> Acute Kidney<br />

Allograft Rejection • Gueler F, Rong S,<br />

Mengel M, Park J-K, Kirsch T, Haller H,<br />

Dumler I, Shushakova N*<br />

091 • a1-antitrypsin Polymerization Studies<br />

using Gas-phase Electrophoretic Mobility<br />

<strong>Molecular</strong> Analysis (GEMMA) • Przygodzki T*,<br />

Mallya M, Phillips RL, Belorgey D, Hägglöf P,<br />

Lomas DA, Ny T<br />

092 • The <strong>Plasminogen</strong> Interaction <strong>of</strong><br />

Antigen 85B Protein from Mycobacterium<br />

Tuberculosis: Role <strong>of</strong> Lys89 • Xolalpa W*,<br />

Vallecillo AJ, Rosales L, Ruiz BH, Espitia C<br />

093 • <strong>Plasminogen</strong> as a Factor in Innate<br />

Immunity • Ahlskog N*, Guo Y, Ny T<br />

094 • The Inflammatory Cytokine Oncostatin<br />

M Induces <strong>Plasminogen</strong> Activator Inhibitor-<br />

1 in Human Vascular Smooth Muscle Cells<br />

in vitro via PI3 kinase <strong>and</strong> MAP-kinase<br />

Dependent Pathways • Demyanets S, Kaun C,<br />

Rychli K, Rega G, Pfaffenberger S, Maurer G,<br />

Huber K, Wojta J*<br />

095 • Crohn’s Disease but not Chronic<br />

Ulcerative Colitis Induces the Expression<br />

<strong>of</strong> PAI-1 in Enteric Neurons • Laerum OD*,<br />

Illemann M, Skarstein A, Helgel<strong>and</strong> L, Øvrebø K,<br />

Danø K, Nielsen BS<br />

096 • The Effect <strong>of</strong> Matrix Metalloprotease<br />

3 Deficiency in Spontaneous Metastasis •<br />

Juncker-Jensen A*, Rømer J, Almholt K<br />

097 • Tumor Cell Expression <strong>of</strong> C4.4A, a<br />

Structural Homologue <strong>of</strong> the Urokinase<br />

Receptor, Correlates with Poor Prognosis<br />

in Non-Small Cell Lung Cancer • Skov BG,<br />

Hansen LV, Ploug M, Pappot H*<br />

098 • Urokinase Receptor Splice Variant<br />

uPAR-del4/5 <strong>and</strong> rab31 mRNA Expression<br />

in Breast Cancer • Magdolen V*, Kotzsch M,<br />

Sieuwerts A, Grosser M, Meye A, Smid M,<br />

Schmitt M, Luther T, Foekens JA<br />

099 • PN-1, a Serine Protease Inhibitor,<br />

Increases MMP-9 Activity in Breast Cancer<br />

Cell Line • Fayard B* <strong>and</strong> Monard D<br />

100 • Cleavage <strong>of</strong> uPAR: Mechanism <strong>and</strong><br />

Prognostic Significance • Høyer-Hansen G*,<br />

Almasi CE, Pappot H<br />

101 • Expression <strong>of</strong> Urokinase Receptor<br />

(uPAR) <strong>and</strong> <strong>Plasminogen</strong> Activator<br />

Inhibitor-1 (PAI-1) in Human Colon Cancer<br />

<strong>and</strong> their Matched Liver Metastases •<br />

Illemann M*, Bird N, Majeed A, Laerum OD,<br />

Lund LR, Danø K, Nielsen BS<br />

102 • A Structural Basis for Differential Cell<br />

Signaling Initiated by PAI-1 <strong>and</strong><br />

PAI-2: Implications for Metastatic Potential •<br />

Croucher D*, Saunders D, Ranson M<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 11


103 • Thrombin Induces Tumor Invasion<br />

through the Induction <strong>and</strong> Association <strong>of</strong><br />

Matrix Metalloproteinase-9 <strong>and</strong> b-1 Integrin<br />

on the Cell Surface • Bruno K*, Radjabi R,<br />

Sawada K, Montag A, Kossiak<strong>of</strong>f A, Lengyel E<br />

104 • Overexpression <strong>of</strong> Protease Nexin-1<br />

mRNA in Oral Squamous Cell Carcinomas •<br />

Gao S, Krogdahl A, Sørensen JA, Dabelsteen E,<br />

Andreasen PA*<br />

105 • The Matrix Metalloprotease (MMP)<br />

Inhibitor Galardin Increases Collagen<br />

Deposition <strong>and</strong> Reduces Spontaneous<br />

Metastasis in the MMTV-PymT Transgenic<br />

Breast Cancer Model • Almholt K*, Lærum OD,<br />

Lund LR, Danø K, Johnsen M, Rømer J<br />

106 • Proteomics <strong>of</strong> uPAR Protein: Protein<br />

Interactions in Cancer Metastasis • Saldanha R,<br />

Molloy M, Xu N, Baker MS*<br />

107 • A New Tagging System for Production<br />

<strong>of</strong> Recombinant Proteins in Drosophila<br />

S2 Cells Using the Third Domain <strong>of</strong> the<br />

Urokinase Receptor • Gårdsvoll H*, Hansen LV,<br />

Jørgensen TJD, Ploug M<br />

108 • Photoaffinity Labeling <strong>of</strong> uPAR with<br />

Cyclic Peptides • Jacobsen B*, Gårdsvoll H,<br />

Barkholt V, Østergaard S, Ploug M<br />

109 • In Vivo Inhibition <strong>of</strong> the Murine<br />

uPA-uPAR Interaction using Monoclonal<br />

Antibodies Raised in uPAR Deficient Mice •<br />

Rasch MG*, Pass J, Jögi A, Rønø B, Gårdsvoll H,<br />

Lund LR, Høyer-Hansen G, Lund IK<br />

110 • RNA Interference for Urokinase-<br />

Targeting Limits Growth <strong>of</strong> Hepatocellular<br />

Carcinoma Xenografts in Nude Mice •<br />

Salvi A*, Arici B, Barlati S, De Petro G<br />

111 • Potent <strong>and</strong> Broad Anti-tumor Activity<br />

<strong>of</strong> an Engineered Matrix Metalloproteinaseactivated<br />

Anthrax Lethal Toxin that Targets<br />

Tumor Vasculature • Liu S, Wang H,<br />

Currie BM, Molinolo A, Leung HJ, Moayeri M,<br />

Alfano RW, Frankel AE, Leppla SH, Bugge TH*<br />

112 • Elucidation <strong>of</strong> the Epitope <strong>of</strong> MA-<br />

31C9, a Non-inhibitory Anti-human PAI-1<br />

Antibody • Meissenheimer LM*, Dewilde M,<br />

Compernolle G, Declerck PJ, Gils A<br />

113 • Residues outside the Epitope Determine<br />

the Function <strong>of</strong> MA-159M12, an Inhibitory<br />

Anti-rat PAI-1 Antibody • Meissenheimer LM*,<br />

Compernolle G, Declerck PJ, Gils A<br />

114 • Conformational Probes <strong>and</strong> Activity<br />

Regulators <strong>of</strong> <strong>Plasminogen</strong> Activator<br />

Inhibitor-1, Isolated from Phage-displayed<br />

Disulphide Bridge-constrained Peptide<br />

Libraries • Dupont DM, Jensen JK, Mathiasen L,<br />

Blouse GE, Wind T, Andreasen PA*<br />

115 • Urokinase-type <strong>Plasminogen</strong> Activatorinhibiting<br />

Cyclic Peptides Demonstrate New<br />

Modalities for Inhibition <strong>of</strong> Serine Proteases •<br />

Andersen LM*, Wind T, Hansen HD, Blouse GE,<br />

Christensen A, Jensen JK, Malmendal A,<br />

Nielsen NC, Andreasen PA<br />

116 • In vivo Treatment with Monoclonal<br />

Antibodies against Mouse Urokinase-type<br />

<strong>Plasminogen</strong> Activator in Cancer Models •<br />

Jögi A*, Lund IK, Høyer-Hansen G, Lund LR,<br />

Danø K, Rømer J<br />

12 X I t h I n t e r n a t i o n a l W o r k s h o p o n


Abstracts for Oral Presentations<br />

i001i<br />

Defining the Native Disulfide Topology in the SMB Domain <strong>of</strong><br />

Human Vitronectin<br />

Li X, Zou G, Yuan W, Lu W*<br />

Institute <strong>of</strong> Human Virology, University <strong>of</strong> Maryl<strong>and</strong> School <strong>of</strong> Medicine, Baltimore, Maryl<strong>and</strong>, USA<br />

Presenting author e-mail: luw@umbi.umd.edu<br />

The N-terminal 44 amino acid residues <strong>of</strong> the human plasma glycoprotein vitronectin, known as<br />

the somatomedin B (SMB) domain, mediates the interaction between vitronectin <strong>and</strong> plasminogen<br />

activator inhibitor 1 (PAI-1) in a variety <strong>of</strong> important biological processes. Several laboratories<br />

have published conflicting reports on the native disulfide topology in the SMB domain with no<br />

consensus reached thus far. Using native chemical ligation <strong>and</strong> orthogonal protection <strong>of</strong> selected<br />

Cys residues, we chemically synthesized three topological analogs <strong>of</strong> SMB with predefined<br />

disulfide connectivities corresponding to those previously published. In addition, we oxidatively<br />

folded a fully reduced SMB in aqueous solution, <strong>and</strong> prepared, by CNBr cleavage, the N-terminal<br />

segment <strong>of</strong> 51 amino acid residues <strong>of</strong> intact vitronectin purified from human blood. Biochemical<br />

<strong>and</strong> functional characterizations allowed us to conclude that (1) only the Cys5-Cys21, Cys9-Cys39,<br />

Cys19-Cys32 <strong>and</strong> Cys25-Cys31 connectivity is present in native vitronectin; (2) only the native<br />

disulfide connectivity is functional; (3) the native disulfide pairings can be readily formed during<br />

spontaneous (oxidative) folding <strong>of</strong> the SMB domain in vitro. Our results unequivocally define the<br />

native disulfide topology in the SMB domain <strong>of</strong> human vitronectin, <strong>and</strong> provide important clues<br />

as to how the controversy arose in the first place.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 13


i002i<br />

The Vitronectin-Binding Site on the Urokinase Receptor<br />

Comprises Residues from Both Domain I <strong>and</strong> the Flanking Interdomain<br />

Linker Region<br />

Gårdsvoll H <strong>and</strong> Ploug M*<br />

Finsen Laboratory, Rigshospitalet, Copenhagen, Denmark<br />

Presenting author e-mail: m-ploug@finsenlab.dk<br />

The urokinase-type plasminogen activator receptor (uPAR) has been implicated as a modulator<br />

<strong>of</strong> several biochemical processes that are active during tumor invasion <strong>and</strong> metastasis e.g.<br />

extracellular proteolysis, cell adhesion <strong>and</strong> cell motility. The structural basis for the high-affinity<br />

interaction between the urokinase-type plasminogen activator (uPA) <strong>and</strong> uPAR, which focuses<br />

cell surface associated plasminogen activation in vivo, is now thoroughly characterized by sitedirected<br />

mutagenesis studies <strong>and</strong> X-ray crystallography. In contrast, the structural basis for the<br />

interaction between uPAR <strong>and</strong> the extracellular matrix protein vitronectin, which is involved<br />

in the regulation <strong>of</strong> cell adhesion <strong>and</strong> motility, remains to be clarified. In the present study, we<br />

have identified the functional epitope on uPAR that is responsible for its interaction with the<br />

full-length, extended form <strong>of</strong> vitronectin using a comprehensive alanine-scanning library <strong>of</strong><br />

purified single-site uPAR mutants (244 positions tested). Interestingly, the 5 residues identified<br />

as ‘’hot spots’’ for vitronectin binding form a contiguous epitope comprising two exposed loops<br />

connecting the central 4-str<strong>and</strong>ed b-sheet in uPAR domain I (Trp32, Arg58 <strong>and</strong> Ile63) as well<br />

as a proximal region <strong>of</strong> the flexible linker peptide connecting uPAR domains I <strong>and</strong> II (Arg91<br />

<strong>and</strong> Tyr92). This binding topology provides the molecular basis for the observation that uPAR<br />

can form a ternary complex with uPA <strong>and</strong> vitronectin. We also show that the affinity for the<br />

small SMB domain <strong>of</strong> vitronectin is increased approximately 4-fold for uPAR-uPA complexes as<br />

compared to unoccupied uPAR.<br />

14 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i003i<br />

uPAR-induced Cell Adhesion <strong>and</strong> Migration:<br />

Vitronectin Provides the Key<br />

Madsen CD* 1 , Sarra Ferraris GM1 , Andolfo A1 , Cunningham O1 1, 2<br />

, Sidenius N<br />

1 The FIRC Institute <strong>of</strong> <strong>Molecular</strong> Oncology (IFOM), Milan, Italy;<br />

2 <strong>Molecular</strong> Genetics Unit, DIBIT, Università Vita-Salute San Raffaele, Milan, Italy<br />

Presenting author e-mail: chris.madsen@ifom-ieo-campus.it<br />

Expression <strong>of</strong> the GPI-anchored membrane receptor uPAR induces pr<strong>of</strong>ound changes in cell<br />

adhesion <strong>and</strong> migration, <strong>and</strong> its expression correlates with the malignant phenotype <strong>of</strong> cancers. To<br />

identify the key molecular interactions essential for uPAR function in these processes; we carried<br />

out a complete functional alanine-scan <strong>of</strong> uPAR in HEK293 cells. Of the 255 mutant receptors<br />

characterized, 34 failed to induce changes in cell adhesion, cell morphology accompanied by<br />

actin-rearrangement <strong>and</strong> focal adhesion turn-over, <strong>and</strong> cell migration. Remarkably, the molecular<br />

defect <strong>of</strong> all <strong>of</strong> these mutants was a specific reduction in integrin-independent cell binding to the<br />

somatomedin-B domain <strong>of</strong> the extracellular matrix component, vitronectin. In order to mimic the<br />

membrane-ECM interaction induced by uPAR-Vn, we generated a GPI-anchored plasminogen<br />

activator inhibitor-1 (PAI-1). This, like uPAR, binds specifically <strong>and</strong> with high affinity to the SMB<br />

domain <strong>of</strong> Vn but shares no other similarity with uPAR. Surprisingly, the chimeric PAI-1/GPI<br />

recapitulated the biological effects <strong>of</strong> uPAR expression. A direct uPAR-Vn interaction is thus both<br />

required <strong>and</strong> sufficient to initiate downstream signalling leading to changes in cell morphology<br />

<strong>and</strong> migration. Together these data demonstrate a novel mechanism by which a cell adhesion<br />

molecule lacking inherent signalling capability evokes complex cellular responses, independently<br />

<strong>of</strong> lateral interactions with signalling receptors, by modulating the contact between the cell <strong>and</strong><br />

the matrix. The importance <strong>of</strong> the uPAR/Vn-interaction was not cell-type specific as all mutants<br />

identified were subsequently confirmed in CHO cells. Finally we have mapped the direct<br />

vitronectin binding epitope (W32, R58, I63, R91, Y92) <strong>of</strong> uPAR.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 15


i004i<br />

Interactions between PAI-1 <strong>and</strong> Vitronectin: Two Proteins,<br />

Two Sites, <strong>and</strong> Two Phases<br />

Schar CR 1 , Jensen JK 2 , Blouse GE 2 , Minor KH 1 , Andreasen PA 2 , Peterson CB* 1<br />

1Department <strong>of</strong> Biochemistry, <strong>Cellular</strong>, <strong>and</strong> <strong>Molecular</strong> <strong>Biology</strong>, University <strong>of</strong> Tennessee, Knoxville,<br />

Tennessee, USA;<br />

2Laboratory <strong>of</strong> <strong>Cellular</strong> Protein Science, Department <strong>of</strong> <strong>Molecular</strong> <strong>and</strong> Structural <strong>Biology</strong>, University <strong>of</strong><br />

Aarhus, Aarhus, Denmark<br />

Presenting author e-mail: cbpeters@utk.edu<br />

We have generated a mutant form <strong>of</strong> vitronectin that lacks the well-characterized N-terminal<br />

somatomedin B domain, known to house the primary high-affinity site for binding <strong>of</strong> PAI-1 to<br />

vitronectin. Residual binding <strong>of</strong> PAI-1 to this deletion mutant <strong>of</strong> vitronectin is observed. Also, we<br />

have used a large battery <strong>of</strong> mutant forms <strong>of</strong> PAI-1 to evaluate the specific interactions required<br />

for binding. With these reagents, the second binding site for vitronectin on PAI-1 was mapped<br />

to a region around helix D rich in charged amino acids. We have used kinetic <strong>and</strong> equilibrium<br />

measurements with surface plasmon resonance to study PAI-1 to full-length <strong>and</strong> the truncated<br />

mutant <strong>of</strong> vitronectin that is missing the somatomedin B domain. Clearly, the interaction <strong>of</strong><br />

vitronectin <strong>and</strong> PAI-1 at the second site is weaker than the primary interaction between the<br />

somatomedin B domain <strong>and</strong> the flexible joint region that lies between helices D, E <strong>and</strong> F on<br />

PAI-1. Most notably, the <strong>of</strong>f rate for binding is much faster, comparable to that for latent PAI-1<br />

dissociation from vitronectin. Interestingly, latent PAI-1 binds nearly as well at the second site<br />

as does active PAI-1, consistent with less dramatic structural changes in the helix D region upon<br />

conversion to the latent structure compared to those that occur with expansion <strong>of</strong> the central beta<br />

sheet that affect the primary binding site interactions with the somatomedin B domain. These<br />

characteristic features <strong>of</strong> binding at the two sites are consistent with FRET experiments <strong>and</strong><br />

stopped-flow fluorescence measurements that reveal biphasic binding between vitronectin <strong>and</strong><br />

PAI-1.<br />

16 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i005i<br />

How Does Vitronectin Accelerate PAI-1’s Protease Inhibition?<br />

Zhou A* <strong>and</strong> Wei Z<br />

Department <strong>of</strong> Haematology, University <strong>of</strong> Cambridge, CIMR, Cambridge, United Kingdom<br />

Presenting author e-mail: awz20@cam.ac.uk<br />

Vitronectin binds <strong>and</strong> stabilises PAI-1’s activity through its somatomedin B (SMB) domain.<br />

Vitronectin can also accelerate PAI-1’s protease (thrombin <strong>and</strong> activated protein C) inhibition<br />

by more than 100-fold. To investigate the mechanism underlying this acceleration, firstly we<br />

prepared SMB domain alone <strong>and</strong> various SMB containing fragments <strong>of</strong> vitronectin. Kinetics<br />

studies showed that these fragments <strong>of</strong> vitronectin, like urea-treated vitronectin, had little effect<br />

on PAI-1’s protease inhibition. Secondly cross-linking experiment with inactive thrombin variant<br />

(S195A) showed that thrombin could only be cross-linked to native VN (not urea-treated VN) in<br />

the presence <strong>of</strong> active PAI-1. Thirdly, mutagenasis studies <strong>of</strong> thrombin showed that substitution<br />

<strong>of</strong> the surface exposed residue Asp100 with Arg attenuated vitronectin PAI-1’s inhibition<br />

<strong>of</strong> thrombin by more than 50%, while other mutations <strong>of</strong> thrombin such as Glu97Arg <strong>and</strong><br />

Asp178Arg etc had no effect. Altogether, these data indicate that native vitronectin has a cryptic<br />

thrombin binding site, which is exposed upon PAI-1 binding (likely masked in urea treated VN)<br />

<strong>and</strong> interacts with thrombin surface near Asp100, <strong>and</strong> vitronectin accelerates PAI-1’s protease<br />

inhibition by bridging PAI-1 <strong>and</strong> thrombin together.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 17


i006i<br />

The Interaction between Tissue-type <strong>Plasminogen</strong> Activator <strong>and</strong><br />

the Low Density Lipoprotein Receptor-Related Protein Induces<br />

<strong>Activation</strong> <strong>of</strong> the NF-kB Pathway during Cerebral Ischemia<br />

Yepes M*, Brotzge XH, Polavarapu R<br />

Department <strong>of</strong> Neurology <strong>and</strong> Center for Neurodegenerative Disease, Emory University School <strong>of</strong><br />

Medicine, Atlanta, Georgia, USA<br />

Presenting author e-mail: myepes@emory.edu<br />

We previously demonstrated that the interaction between tPA <strong>and</strong> the low density lipoprotein<br />

receptor-related protein (LRP) following middle cerebral artery occlusion (MCAO) increases<br />

the permeability <strong>of</strong> the blood brain barrier (BBB). Here we studied the relation between tPA <strong>and</strong><br />

NF-kB activation following MCAO. Wild-type (WT), tPA (tPA–/–), <strong>and</strong> plasminogen (Plg–/–)<br />

deficient mice underwent MCAO <strong>and</strong> analysis <strong>of</strong> NF-kB activation by immunohistochemistry,<br />

Western blot (p65-phosphorylation) <strong>and</strong> electrophoretic mobility shift assay (EMSA). We observed<br />

a rapid activation <strong>of</strong> NF-kB in WT <strong>and</strong> Plg–/– mice that was abolished in tPA–/– animals. The<br />

effect <strong>of</strong> MCAO on LRP expression was studied by immunohistochemistry <strong>and</strong> quantitative<br />

real-time PCR (qRT-PCR). We found that MCAO induces a rapid increase in LRP expression<br />

in WT mice that is significantly attenuated in tPA–/– mice. Treatment <strong>of</strong> WT mice with either<br />

the receptor associated protein (RAP) or anti-LRP antibodies inhibited MCAO-induced NF-kB<br />

activation. The intracerebral injection <strong>of</strong> tPA into tPA–/– mice following MCAO resulted in NF-kB<br />

activation that was not observed when tPA was co-administered with either RAP or LRP-blocking<br />

antibodies. Immunohistochemistry <strong>and</strong> qRT-PCR demonstrated an increase in the expression <strong>of</strong><br />

inducible nitric oxide synthase (NF-kB-dependent gene) in the ischemic area in WT mice that<br />

was attenuated in tPA–/– mice <strong>and</strong> in WT animals treated with RAP or anti-LRP antibodies.<br />

We conclude that during cerebral ischemia tPA has a plasminogen-independent ‘’cytokine-like’’<br />

function <strong>and</strong> that the interaction between tPA <strong>and</strong> LRP results in NF-kB pathway activation <strong>and</strong><br />

induction <strong>of</strong> NF-kB-dependent genes with known effect on cell death <strong>and</strong> BBB permeability.<br />

18 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i007i<br />

Unique Secretory Dynamics <strong>of</strong> Tissue <strong>Plasminogen</strong> Activator (tPA)<br />

Is Beneficial to Maintain Fibrinolytic Activity on Cell Surface<br />

Suzuki Y*, Ihara H, Mogami H, Urano T<br />

Department <strong>of</strong> Physiology, Hamamatsu University School <strong>of</strong> Medicine, Hamamatsu, Japan<br />

Presenting author e-mail: seigan@hama-med.ac.jp<br />

Introduction: Vascular endothelial cells (VECs) express <strong>and</strong> secrete various anti-thrombotic<br />

molecules to keep the vascular patency. tPA, the primary fibrinolytic enzyme in vasculature, is<br />

one <strong>of</strong> them. After secretion as an active form, tPA initiates fibrinolysis both on VECs <strong>and</strong> in blood<br />

where its specific inhibitor <strong>of</strong> PAI-1 exists. Here, we analyzed the dynamics <strong>of</strong> tPA secretion from<br />

the containing granules <strong>and</strong> its modulation by PAI-1 using total internal reflection fluorescence<br />

microscopy (TIRF-M). Method: An established cell-line <strong>of</strong> VECs was cultured <strong>and</strong> transfected<br />

with green fluorescent protein (GFP)-tagged either wild type tPA, tPA (S478A)(Ser478 at active<br />

center is replaced by Ala, no ability to complex with PAI-1) or tPA (catalytic domain; CD)(deleted<br />

in finger-, EGF-like-, kringle1- <strong>and</strong> 2- domains). The exocytotic dynamics <strong>of</strong> these tPAs-GFP near<br />

the plasma membrane (PM) were analyzed by TIRF-M. Results: (1) tPA-GFP showed unique<br />

dynamics <strong>of</strong> slow disappearance from PM after opening <strong>of</strong> its containing granules (fluorescence<br />

half life: TF1/2=10sec). (2) Supplemented PAI-1 shortened TF1/2 <strong>and</strong> increased tPA-PAI-1<br />

complex but not free tPA in supernatant. TF1/2 <strong>of</strong> tPA(S478A)-GFP was slower than tPA-GFP,<br />

indicating that the complex formation with PAI-1 is essential for rapid dissociation <strong>of</strong> tPA from<br />

the opened granular membrane. (3) TF1/2 <strong>of</strong> tPA(CD)-GFP was faster than tPA-GFP, indicating<br />

that FEK domains are responsible for slow disappearance. Conclusion: tPA has unique, slow<br />

secretory dynamics which is beneficial to maintain fibrinolytic activity on VECs. PAI-1 facilitates<br />

tPA dissociation <strong>and</strong> suppresses fibrinolytic activity on VECs.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 19


i008i<br />

The Macrophage Low-density Lipoprotein-Receptor-Related<br />

Protein (LRP) Modulates Murine Lipoprotein Metabolism<br />

Lillis AP* 1,2 , Mikhailenko I 1 , Robinson S 1 , Migliorini M 1 , Battey F 1 , Pizzo SV 2 , Strickl<strong>and</strong> DK 1<br />

1Center for Vascular <strong>and</strong> Inflammatory Diseases, University <strong>of</strong> Maryl<strong>and</strong> School <strong>of</strong> Medicine, Baltimore,<br />

Maryl<strong>and</strong>, USA;<br />

2 Department <strong>of</strong> Pathology, Duke University Medical Center, Durham, North Carolina, USA<br />

Presenting author e-mail: alillis@som.umaryl<strong>and</strong>.edu<br />

Very low-density lipoproteins (VLDL) <strong>and</strong> chylomicrons transport cholesterol <strong>and</strong> triglycerides<br />

(TG) to muscle, adipose <strong>and</strong> other extra-hepatic tissues. Lipoprotein lipase selectively removes<br />

<strong>and</strong> hydrolyzes TGs, transferring free fatty acids to these tissues. The resulting remnant<br />

lipoproteins bind to two members <strong>of</strong> the low-density lipoprotein receptor (LDLR) family, LDLR<br />

itself <strong>and</strong> the LDLR-related protein (LRP). LRP is a large endocytic receptor that recognizes more<br />

than 30 different lig<strong>and</strong>s. Accumulating evidence supports a role for hepatic LRP in the clearance<br />

<strong>of</strong> remnant lipoproteins. However, LRP is also abundant in resident liver macrophages (Kupffer<br />

cells) which reside in the sinusoids <strong>of</strong> the liver in close proximity to the space between endothelial<br />

cells <strong>and</strong> hepatocytes where lipoprotein remnants are sequestered <strong>and</strong> modified before being<br />

recognized by LRP <strong>and</strong> the LDL receptor. To investigate a possible role <strong>of</strong> macrophage LRP in<br />

remnant lipoprotein metabolism, we developed tissue-specific knockouts using the LoxP/Cremediated<br />

recombination system on mice genetically deficient in the LDLR. By crossing mice<br />

carrying loxP-tagged LRP with a LysMCre transgenic mouse line expressing the Cre recombinase<br />

at the mouse M lysozyme locus, we have generated LDLR-deficient mice in which the LRP gene<br />

is also selectively <strong>and</strong> efficiently deleted in macrophages (macLRP- mice). After placing macLRP-<br />

mice (<strong>and</strong> their sibling controls which are deficient in LDLR but express LRP normally) on a high<br />

fat, high cholesterol (Western) diet for 3 weeks, we observed elevated serum triglycerides (5.0 vs<br />

2.8 mg/ml, p


i009i<br />

Annexin 2 Mediates <strong>Plasminogen</strong>-Dependent Recruitment <strong>of</strong><br />

Neovascular Mural Cells in Lymphoma Angiogenesis<br />

Ling Q 1 , Ruan J 2 , Yan L 1 , Sui G-Z 1 , Deora AB 1 , Church S 2 , Cohen-Gould L 1 , Rafii S 2 , Lyden D 3 ,<br />

Hajjar KA* 1<br />

Departments <strong>of</strong> 1 Cell <strong>and</strong> Developmental <strong>Biology</strong>, 2 Medicine, 3 Pediatrics, Weill Cornell Medical College,<br />

New York, New York, USA<br />

Presenting author e-mail: khajjar@med.cornell.edu<br />

Malignant tumor progression depends upon the development <strong>of</strong> tumor-associated blood vessels,<br />

either via co-option <strong>of</strong> nearby host vascular cells, or through recruitment <strong>of</strong> marrow-derived<br />

progenitor cells. Annexin 2 (A2) is a cell surface co-receptor for plasminogen (Plg) <strong>and</strong> tPA that<br />

augments the catalytic efficiency <strong>of</strong> Plg activation. We showed previously that A2–/– mice display<br />

impaired growth factor-induced postnatal angiogenesis in the corneal pocket, oxygen-induced<br />

retinopathy, <strong>and</strong> Matrigel plug assays. Here, we examined neovascularization <strong>of</strong> experimental<br />

lymphoma in mice with fibrinolytic deficiency states. For EL4, a T cell lymphoma, <strong>and</strong> B6RV2, a<br />

B cell lymphpma, tumor growth was rapid in wildtype <strong>and</strong> tPA–/– mice, but severely retarded<br />

in both A2–/– <strong>and</strong> Plg–/– mice. Immunohistochemical staining <strong>of</strong> EL4 tissue on day 8 revealed<br />

reduced vascular density, frequent intravascular fibrin thrombi, <strong>and</strong> dilated microvessels in<br />

A2–/–, but not A2+/+, mice. Electron microscopy <strong>and</strong> immun<strong>of</strong>luorescence revealed a paucity<br />

<strong>of</strong> pericytes <strong>and</strong> secondary dropout <strong>of</strong> endothelial cells within A2–/– tumor microvessels. Flow<br />

cytometric analysis <strong>of</strong> circulating marrow-derived progenitor cells showed reduced populations<br />

<strong>of</strong> VEGFR1+/CD11b+ hematopoietic precursor cells in A2–/– versus A2+/+ tumor-bearing mice.<br />

In lethally irradiated A2–/– mice, tumor growth was rescued completely upon engraftment<br />

with A2+/+ marrow. Transplantation <strong>of</strong> green fluorescent protein (GFP)-positive bone marrow<br />

led to the recruitment <strong>of</strong> abundant GFP+/CD11b+/CD68+ cells to locations surrounding<br />

tumor neovessels. In addition, transplantation <strong>of</strong> normal marrow restored the investment <strong>of</strong><br />

tumor microvessels with a-smooth muscle actin-positive pericytes. These data indicate that A2<br />

contributes critically to tumor angiogenesis in experimental lymphoma, by enabling recruitment<br />

<strong>of</strong> A2+ myelomonocytic cells that instruct pericyte recruitment <strong>and</strong> neovascular stabilization.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 21


i010i<br />

<strong>Activation</strong> <strong>of</strong> Latent PDGF-CC by Tissue <strong>Plasminogen</strong> Activator<br />

Impairs Blood Brain Barrier Integrity during Ischemic Stroke<br />

Su EJ 1 , Fredriksson L 2 , Geyer M 1 , Folestad E 2 , Cale J 1 , Mann K 1 , Gao Y 3 , Pietras K 2 , Andreé J 4 , Yepes M 5 ,<br />

Strickl<strong>and</strong> D 3 , Betsholtz C 4 , Eriksson U 2 , Lawrence DA* 1<br />

1Department <strong>of</strong> Internal Medicine, Division <strong>of</strong> Cardiovascular Medicine, University <strong>of</strong> Michigan<br />

Medical School, Ann Arbor, Michigan, USA;<br />

2 Ludwig Institute for Cancer Research, Stockholm Branch, Karolinska Institutet, Stockholm, Sweden;<br />

3Center for Vascular <strong>and</strong> Inflammatory Disease <strong>and</strong> Departments <strong>of</strong> Surgery <strong>and</strong> Physiology, School <strong>of</strong><br />

Medicine, University <strong>of</strong> Maryl<strong>and</strong>, Baltimore, Maryl<strong>and</strong>, USA;<br />

4Laboratory <strong>of</strong> Vascular <strong>Biology</strong>, Division <strong>of</strong> Matrix <strong>Biology</strong>, Department <strong>of</strong> Medical Biochemistry <strong>and</strong><br />

Biophysics, Karolinska Institutet, Stockholm, Sweden;<br />

5Department <strong>of</strong> Neurology <strong>and</strong> Center for Neurodegenerative Disease, Emory University School <strong>of</strong><br />

Medicine, Atlanta, Georgia, USA<br />

Presenting author e-mail: dlawrenc@umich.edu<br />

The current treatment for ischemic stroke, intravenous tPA, only benefits a limited number <strong>of</strong><br />

patients. This is due in part to the requirement that tPA be administered within 3 hours <strong>of</strong> the<br />

onset <strong>of</strong> symptoms. The reasons for this time constraint are not known but may be due to the<br />

unique activities that tPA has in the brain beyond its well established role as a fibrinolytic enzyme.<br />

For example, tPA has been shown to play a role in regulating cerebrovascular permeability<br />

after stroke; however, the specific substrate for tPA in brain is not known. The recent discovery<br />

that tPA cleaves latent PDGF-CC lead us to hypothesize that tPA may mediate cerebrovascular<br />

permeability via activation <strong>of</strong> latent PDGF-CC. To test this possibility, active PDGF-CC was<br />

injected directly into the CSF in the absence <strong>of</strong> ischemia <strong>and</strong> the extravasation <strong>of</strong> Evans Blue<br />

dye in the brain was examined. These studies showed a significant increase in Evans Blue<br />

extravasation after active PDGF-CC injection. This effect was similar to that seen with tPA<br />

injection, suggesting that tPA <strong>and</strong> PDGF-CC may be acting on the same pathway. To test this<br />

neutralizing antibodies against PDGF-CC were co-injected with tPA <strong>and</strong> were found to block<br />

the effect <strong>of</strong> tPA. To evaluate whether PDGF-CC plays a role in regulating cerebrovascular<br />

permeability during stroke, the PDGF receptor inhibitor Gleevec was used in a photothrombotic<br />

stroke model. Stroke is induced specifically in the middle cerebral artery by the local activation<br />

<strong>of</strong> intravenous Rose Bengal with a cold laser. Mice treated with Gleevec showed significant<br />

reductions in Evans Blue extravasation at 24-hours <strong>and</strong> stroke volumes at 72-hours, indicating a<br />

strong effect <strong>of</strong> PDGF signaling on cerebrovascular integrity. Taken together, these data suggest<br />

that PDGF-CC is a downstream substrate <strong>of</strong> tPA in the CNS. These studies may provide important<br />

insights into new therapeutic strategies for treating stroke patients.<br />

22 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i011i<br />

<strong>Plasminogen</strong> Activator Inhibitor-1 Gene Regulation:<br />

Cross Talk between Hypoxia <strong>and</strong> Insulin Signalling<br />

Flugel D <strong>and</strong> Kietzmann T*<br />

Faculty <strong>of</strong> Chemistry, Department Biochemistry, University <strong>of</strong> Kaiserslautern, Kaiserslautern, Germany<br />

Presenting author e-mail: tkietzm@gwdg.de<br />

A number <strong>of</strong> pathological conditions like cancer, diabetes or the metabolic syndrome are<br />

associated with enhanced plasminogen activator inhibitor-1 (PAI-1) levels. In addition, these<br />

diseases are <strong>of</strong>ten combined with hypoxia <strong>and</strong> an impaired insulin signalling. Therefore we asked<br />

whether hypoxia <strong>and</strong> insulin can exert direct effects on PAI-1 expression. We found that hypoxia<br />

induced PAI-1 expression via the hypoxia-inducible transcription factor-1 (HIF-1) which binds<br />

to hypoxia responsive elements in the PAI-1 promoter. HIF-1 is a heterodimer from which HIF-<br />

1alpha becomes hydroxylated under normoxia. This enables binding <strong>of</strong> the von Hippel-Lindau<br />

(VHL) protein which targets HIF-1alpha for proteasomal degradation.<br />

Interestingly, HIF-1a can also respond to non-hypoxic stimuli like insulin which acts via<br />

phosphatidylinositol 3-kinase <strong>and</strong> protein kinase B. Although PKB induces HIF-1alpha<br />

stabilization, HIF-1a is not a direct substrate for PKB/Akt. Therefore, we aimed to investigate<br />

whether glycogen synthase kinase-3 (GSK3) may have an impact on the VHL-dependent HIF-<br />

1alpha degradation.<br />

We found that inhibition <strong>of</strong> GSK3 with LiCl <strong>and</strong> insulin as well as depletion with siRNA induced<br />

HIF-1alpha <strong>and</strong> PAI-1 expression whereas overexpression <strong>of</strong> GSK3alpha <strong>and</strong> GSK3beta reduced it.<br />

These effects were mediated posttranslationally via the oxygen-dependent degradation domain <strong>of</strong><br />

HIF-1alpha. Mutation <strong>of</strong> the proline residues critical for the VHL-dependent degradation as well<br />

as usage <strong>of</strong> VHL-deficient cells did not prevent GSK3-mediated ubiquitylation <strong>and</strong> degradation <strong>of</strong><br />

HIF-1alpha. However, inhibition <strong>of</strong> the proteasome by MG132 partially reversed the GSK3 effects<br />

indicating that GSK3 could target HIF-1alpha to the proteasome by phosphorylation. Further,<br />

we identified three putative target sites for GSK3 within HIF-1alpha <strong>and</strong> mutation <strong>of</strong> these sites<br />

increased HIF-1alpha transactivity, protein stability <strong>and</strong> PAI-1 expression.<br />

Thus, the present data show that hypoxia <strong>and</strong> insulin signalling merge at HIF-1alpha which<br />

appears to have a prominent role for the regulation <strong>of</strong> PAI-1 expression.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 23


i012i<br />

Role <strong>of</strong> Rho GTPases <strong>and</strong> p38 MAP Kinase in the Regulation <strong>of</strong><br />

t-PA <strong>and</strong> PAI-1 Expression in Cultured Human Endothelial Cells<br />

Fish RJ*, Dunoyer-Geindre S, Kruith<strong>of</strong> EKO<br />

Service <strong>of</strong> Angiology <strong>and</strong> Haemostasis, Department <strong>of</strong> Internal Medicine, Geneva University Hospital<br />

<strong>and</strong> Medical School, Geneva, Switzerl<strong>and</strong><br />

Presenting author e-mail: Richard.Fish@medecine.unige.ch<br />

HMG-CoA reductase inhibitors, or statins, are used as cholesterol-lowering drugs for<br />

cardiovascular disease (CVD). Clinical evidence suggests that statins are also beneficial for CVD<br />

patients independently from their effects on cholesterol levels. Statins increase endothelial cell<br />

expression <strong>of</strong> t-PA while lowering levels <strong>of</strong> PAI-1. This could increase the fibrinolytic potential<br />

<strong>of</strong> the endothelium <strong>and</strong> be beneficial in CVD. We investigated the mechanism <strong>of</strong> statin-induced<br />

changes in t-PA <strong>and</strong> PAI-1 expression in HUVEC. t-PA <strong>and</strong> PAI-1 mRNA were measured by<br />

quantitative RT-PCR. A time- <strong>and</strong> dose-dependent increase in t-PA, <strong>and</strong> decrease in PAI-1,<br />

were measured in fluvastatin-treated HUVEC. These changes were reversed by mevalonate<br />

or geranylgeranyl pyrophosphate, <strong>and</strong> mimicked by a geranylgeranyl transferase inhibitor -<br />

demonstrating that the statins effects were mediated by inhibition <strong>of</strong> protein geranylgeranylation.<br />

Rho family GTPases are geranylgeranylated <strong>and</strong> therefore potential targets for the effects <strong>of</strong><br />

fluvastatin. Dominant negative (DN) RhoA, Rac1 <strong>and</strong> Cdc42 were expressed in HUVEC. A<br />

minor increase in t-PA expression was measured in cells expressing DNRhoA, while DNRac1<br />

or DNCdc42 expression each lead to a three-fold increase. PAI-1 expression decreased in cells<br />

expressing DNRhoA or DNRac1, but not in cells expressing DNCdc42. Fluvastatin-induced t-PA<br />

expression is most likely mediated by p38 MAP kinase (p38) because the p38 inhibitor, SB202190,<br />

reversed fluvastatin-induced t-PA expression but not the decrease in PAI-1. Our results show<br />

a role for p38 in statin-induced t-PA, with geranylgeranylated Rho GTPases as potential statin<br />

targets. Endothelial t-PA <strong>and</strong> PAI-1 expression appear to be differentially regulated by Rho<br />

GTPases.<br />

24 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i013i<br />

Urokinase Receptor/Alpha5Beta1 Integrin Interaction <strong>and</strong><br />

Signaling in Cancer Cells<br />

Wei Y*, Tang CH, Kim Y, Robillard L, Kugler MC, Hill M, Brumwell A, Chapman HA<br />

Pulmonary <strong>and</strong> Critical Care Division, Department <strong>of</strong> Medicine, Cardiovascular Research Institute,<br />

University <strong>of</strong> California, San Francisco, California, USA<br />

Presenting author e-mail: ying.wei@ucsf.edu<br />

The urokinase receptor (uPAR) is commonly upregulated in tumor cells <strong>and</strong> thought to promote<br />

cancer invasion <strong>and</strong> metastasis. Our previous studies show that uPAR binds fibronectin (Fn)<br />

receptor a5b1 integrin, the complex shifting the integrin binding site on Fn from the canonical<br />

RGD motif to a nearby Fn heparin binding domain (HepII). We hypothesized that uPAR<br />

interactions with a5b1 modulate Fn signaling <strong>and</strong> promote tumor progression in vivo. To study the<br />

effects <strong>of</strong> uPAR/a5b1 interactions on signaling, endogenous uPAR was stably silenced by RNAi in<br />

malignant tumor cell lines HT1080 <strong>and</strong> H1299 <strong>and</strong> reconstituted with WT uPAR or a uPAR point<br />

mutant (H249A) discovered to be defective in a5b1-binding. Fn attachment initiated src/FAK,<br />

Rac1, <strong>and</strong> ERK activation followed by MMP-9 expression. In both uPAR knockdown <strong>and</strong> H249A<br />

uPAR bearing tumor cells Fn adhesion initiated src/FAK activation but was unable to activate<br />

Rac1 or ERK or induce MMP-9. Thus MMP-9 induction by Fn depends on two signals: one from<br />

a5b1 leading to active src/FAK <strong>and</strong> one from a5b1/uPAR leading to active Rac1. Both are required<br />

from ERK activation <strong>and</strong> MMP-9 induction. In vivo, knockdown <strong>of</strong> uPAR in tumor cells injected<br />

intravenously or into the lung orthotopically resulted in markedly reduced tumor incidence <strong>and</strong><br />

size. UPAR knockdown cells maintained low levels <strong>of</strong> MMP-9 after at least two days in the mouse<br />

lungs. Collectively, these data demonstrate that uPAR/a5b1 complexes are required for MMP-9<br />

induction through an ERK-dependent <strong>and</strong> that this signaling pathway may be relevant to tumor<br />

development in vivo.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 25


i014i<br />

Activated Human Neutrophils Rapidly Release the Chemotactically<br />

Active D2D3 Form <strong>of</strong> the Urokinase-type <strong>Plasminogen</strong> Activator<br />

Receptor (uPAR/CD87)<br />

Pliyev BK* 1,2 <strong>and</strong> Tkachuk VA 1,2<br />

1Department <strong>of</strong> Biological <strong>and</strong> Medical Chemistry, School <strong>of</strong> Basic Medicine, Moscow State University,<br />

Moscow, Russia;<br />

2<strong>Molecular</strong> Endocrinology Laboratory, Institute <strong>of</strong> Experimental Cardiology, Cardiology Research<br />

Center, Moscow, Russia<br />

Presenting author e-mail: bpliyev@cardio.ru<br />

The urokinase-type plasminogen activator receptor (uPAR/CD87) exists both in cell-bound <strong>and</strong><br />

soluble forms. Soluble uPAR (suPAR) is readily detected in blood <strong>and</strong> is markedly increased<br />

during inflammation <strong>and</strong> cancer. Neutrophils contain extensive intracellular pools <strong>of</strong> uPAR that<br />

are translocated to the plasma membrane upon activation. In the present study, we investigated<br />

the ability <strong>of</strong> human neutrophils to shed uPAR from cell surface following activation <strong>and</strong><br />

addressed the possible involvement <strong>of</strong> the released receptor in inflammatory response. We<br />

first observed that resting neutrophils spontaneously release suPAR. This release was strongly<br />

<strong>and</strong> rapidly (within minutes) enhanced by calcium ionophore ionomycin <strong>and</strong> to a lesser extent<br />

when cells were primed with TNF-alpha or LPS <strong>and</strong> then stimulated with fMLP or IL-8. We<br />

demonstrate that suPAR is produced by resting <strong>and</strong> activated neutrophils predominantly as a<br />

truncated form devoid <strong>of</strong> N-terminal D1 domain (D2D3 form) that lacks GPI anchor. Migration<br />

<strong>of</strong> human embryonic kidney (HEK) 293 cells stably transfected with the human formyl peptide<br />

receptor FPRL1 towards the supernatants harvested from activated neutrophils was significantly<br />

diminished when suPAR was immunodepleted from the supernatants. We conclude that activated<br />

neutrophils release the chemotactically active D2D3 form <strong>of</strong> suPAR that acts as a lig<strong>and</strong> <strong>of</strong><br />

FPRL1. The release <strong>of</strong> suPAR by activated cells was significanty but not completely inhibited by<br />

alpha 1-antichymotrypsin, a specific inhibitor <strong>of</strong> cathepsin G, indicating that cathepsin G acts as<br />

a shedding protease for membrane-bound uPAR. Neutrophils isolated from synovial fluids <strong>of</strong><br />

rheumatoid arthritis patients released significantly (p < 0.01) higher amounts <strong>of</strong> suPAR compared<br />

with cells isolated from pared peripheral blood samples suggesting that the release <strong>of</strong> suPAR by<br />

neutrophils is increased in vivo in the sites <strong>of</strong> acute inflammation. We suggest that production <strong>of</strong><br />

the chemotactically active D2D3 form <strong>of</strong> suPAR by activated human neutrophils in vivo could<br />

contribute to the recruitment <strong>of</strong> monocytes <strong>and</strong> other formyl peptide receptors-expressing cells to<br />

the sites <strong>of</strong> acute inflammation where neutrophils accumulation <strong>and</strong> activation occur.<br />

26 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i015i<br />

PDGF-DD Bioavailability Is Regulated by the uPA/uPAR System:<br />

Implications for Tumor Growth<br />

Ehnman M*, Li H, Fredriksson L, Eriksson U<br />

Ludwig Institute for Cancer Research, Stockholm Branch, Karolinska Institutet, Stockholm, Sweden<br />

Presenting author e-mail: monika.ehnman@licr.ki.se<br />

Members <strong>of</strong> the Platelet-derived growth factor (PDGF) family are involved in many physiological<br />

<strong>and</strong> pathological events such as embryonic development, blood vessel maturation, fibrotic<br />

disease <strong>and</strong> cancer. In contrast to the classical PDGFs, the novel <strong>and</strong> closely related PDGF-CC<br />

<strong>and</strong> PDGF-DD are latent factors that need to be extracellularly processed before they can initiate<br />

PDGF receptor activation. We have previously identified tissue plasminogen activator (tPA) as<br />

an activator <strong>of</strong> PDGF-CC <strong>and</strong> we are now elucidating the biological relevance <strong>of</strong> the urokinase<br />

plasminogen activator (uPA) <strong>and</strong> its receptor for the proteolytic activation <strong>of</strong> PDGF-DD.<br />

Both tPA <strong>and</strong> uPA are plasminogen activators accounting for the generation <strong>of</strong> plasmin in<br />

different tissues, including blood, thereby initializing a proteolytic cascade leading to breakdown<br />

<strong>of</strong> extracellular matrix components, activation <strong>of</strong> other proteases <strong>and</strong> growth factors. The uPA/<br />

uPAR system has been extensively studied during the years <strong>and</strong> the components have been<br />

shown to play a role as prognostic markers in various cancers <strong>and</strong> in the metastatic process.<br />

Here we have elucidated the structural requirements for uPA-mediated activation <strong>of</strong> PDGF-DD.<br />

To further explore if this activation process may play a role in NIH/3T3 cell transformation <strong>and</strong><br />

subsequent tumor growth in nude mice, we generated stable NIH/3T3 cell lines expressing either<br />

activated PDGF-DD or latent PDGF-DD <strong>and</strong> compared their in vitro growth capacity <strong>and</strong> in vivo<br />

tumorigenicity. Our data suggest that regulated PDGF-DD activation is important for NIH/3T3<br />

tumor development in vivo <strong>and</strong> indicate that the uPA/uPAR system may be <strong>of</strong> importance for<br />

localized <strong>and</strong> regulated PDGF-DD activation.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 27


i016i<br />

TGF-b1-Induced <strong>Plasminogen</strong> Activator Inhibitor-<br />

1 Expression in Vascular Smooth Muscle Cells Requires<br />

Cooperative Rho/ROCK <strong>and</strong> EGFR Signaling<br />

Higgins PJ* <strong>and</strong> Samarakoon R<br />

Center for Cell <strong>Biology</strong> & Cancer Research, Albany Medical College, Albany, New York, USA<br />

Presenting author e-mail: higginp@mail.amc.edu<br />

TGF-b1 <strong>and</strong> its target gene encoding plasminogen activator inhibitor-1 (PAI-1) are major causative<br />

factors in the pathophysiology <strong>of</strong> arteriosclerosis <strong>and</strong> perivascular fibrosis. The definition <strong>of</strong> TGFb1-activated<br />

signaling networks that impact PAI-1 transcription in vascular smooth muscle cells<br />

(VSMC) may identify novel, therapeutically-relevant, opportunities to manage PAI-1-associated<br />

cardiovascular disease. TGF-b1-induced PAI-1 expression in VSMC required cooperative EGFR/<br />

MEK-ERK <strong>and</strong> Rho/ROCK signaling. Transient transfection <strong>of</strong> a dominant negative RhoA (DN-<br />

RhoA) expression construct or pretreatment <strong>of</strong> VSMC with C3 transferase (a Rho inhibitor) <strong>and</strong><br />

Y-27632 (an inhibitor <strong>of</strong> p160ROCK, a downstream effector <strong>of</strong> Rho) dramatically attenuated the<br />

TGF-b1-initiated PAI-1 inductive response. Genetic EGFR1 deficiency or pharmacologic inhibition<br />

<strong>of</strong> EGFR activity (with AG1478) virtually ablated TGF-b1-stimulated ERK1/2 activation as well<br />

as PAI-1 expression but not SMAD2 phosphorylation. Interference with Rho/ROCK signaling,<br />

in contrast, prevented SMAD2 activation <strong>and</strong> nuclear accumulation. Infection <strong>of</strong> VSMC with<br />

an adenovirus encoding a mutant Y845F-EGFR effectively decreased TGF-b1-induced PAI-1<br />

expression implicating the pp60c-src phosphorylation site (Y845) <strong>of</strong> the EGFR in the inductive<br />

response. This is consistent with previous findings that pp60c-src is activated by TGF-b1 in<br />

VSMC <strong>and</strong> that dominant negative pp60c-src (DN-Src) expression vectors significantly block<br />

TGF-b1 induced PAI-1 expression. SMAD2 activation, moreover, is not sufficient to induce PAI-<br />

1 expression by TGF-b1 in the absence <strong>of</strong> EGFR signaling both in VSMC <strong>and</strong> mouse embryonic<br />

fibroblasts. Thus, two distinct but cooperative pathways involving EGFR/MEK-ERK signaling<br />

<strong>and</strong> Rho-dependent SMAD2 activation are required for TGF-b1-induced PAI-1 expression in<br />

VSMC.<br />

28 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i017i<br />

A Novel Neuronal Death Pathway Triggered by<br />

Excess Tissue <strong>Plasminogen</strong> Activator<br />

Li J* 1 , Snyder EY 2 , Sidman RL 1<br />

1Department <strong>of</strong> Neurology, Beth Israel Deaconess Medical Center, Harvard Medical School, Boston,<br />

Massachusetts, USA;<br />

2 Stem Cell <strong>and</strong> Regeneration Program, The Burnham Institute for Medical Research, La Jolla, California,<br />

USA<br />

Presenting author e-mail: jli7@caregroup.harvard.edu<br />

We have explored molecular features <strong>of</strong> the cell death pathway in nervous (nr) mutant mice<br />

that show mitochondrial abnormalities <strong>and</strong> dendritic growth retardation in cerebellar Purkinje<br />

neurons (PNs) in postnatal weeks 2-3, <strong>and</strong> massive PN degeneration <strong>and</strong> motor incoordination<br />

thereafter. We found a mutation-induced dramatic 10-fold excess <strong>of</strong> tissue plasminogen activator<br />

(tPA) in nr cerebellum, affecting two types <strong>of</strong> downstream targets: voltage-dependent anion<br />

channels (VDACs) <strong>and</strong> neurotrophins. VDACs affect synaptic communication <strong>and</strong> apoptotic<br />

cell death, as well as mitochondrial metabolite fluxes. The tPA/plasmin system contains five<br />

‘’kringle’’ binding segments, <strong>of</strong> which kringle 5 can specifically bind brain mitochondrial VDACs<br />

to induce partial closure <strong>of</strong> these channels <strong>and</strong> interfere with mitochondria-related regulation <strong>of</strong><br />

intracellular calcium <strong>and</strong> pH. Neurotrophins participate in regulation <strong>of</strong> neuronal morphogenesis<br />

<strong>and</strong> maintenance. Degradation <strong>of</strong> particular neurotrophins by tPA/plasmin may account for<br />

tPA-mediated decreases in synaptic organization <strong>and</strong> neuronal survival. We have established<br />

that neural stem cells (NSCs) transplanted into neonatal nr cerebellum rescued PNs from cell<br />

death weeks later by restoring molecular homeostasis <strong>of</strong> tPA <strong>and</strong> its downstream targets. Two<br />

weeks after NSC injection, <strong>and</strong> a week before onset <strong>of</strong> PN degeneration, correction <strong>of</strong> tPA was<br />

accompanied by 1) normal amount <strong>and</strong> distribution <strong>of</strong> VDACs <strong>and</strong> normal mitochondrial shape,<br />

<strong>and</strong> 2) neurotrophin normalcy <strong>and</strong> PN dendritic growth. Likewise, NSCs contacting cerebellar<br />

slices in vitro rectified tPA levels <strong>and</strong> rescued host PNs, as did pharmacological decrease <strong>of</strong> tPA to<br />

normal but not subnormal levels. Neurotoxic action <strong>of</strong> tPA is open to further study in this system.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 29


i018i<br />

The Inhibitor <strong>of</strong> Serine Proteases Protease Nexin-1 <strong>and</strong> its Receptor<br />

LRP Modulate SHH Signalling during Cerebellar Development<br />

Vaillant C 1 , Michos O 2 , Orolicki S 1 , Brellier F 1 , Taieb S 1 , Moreno E 1 , Té H 1 , Zeller R 2 , Monard D* 1<br />

1 Friedrich Miescher Institute for Biomedical Research, Basel, Switzerl<strong>and</strong>;<br />

2 Developmental Genetics, DKBW Center for Biomedicine, University <strong>of</strong> Basel Medical School, Basel,<br />

Switzerl<strong>and</strong><br />

Presenting author e-mail: denis.monard@fmi.ch<br />

Development <strong>of</strong> the postnatal cerebellum relies on tight regulation <strong>of</strong> the cell number by<br />

morphogens that control the balance between cell proliferation, survival <strong>and</strong> differentiation. Here<br />

we analyze the role <strong>of</strong> the serine protease inhibitor Protease Nexin-1 (PN-1) in the proliferation<br />

<strong>and</strong> differentiation <strong>of</strong> Cerebellar Granular Neuron Precursors (CGNPs) via modulation <strong>of</strong> their<br />

main mitogenic factor, Sonic Hedgehog (SHH). Our studies show that PN-1 interacts with the<br />

Low-density lipoprotein receptor-Related Proteins (LRPs) to antagonize SHH-induced CGNP<br />

proliferation <strong>and</strong> inhibits the activity <strong>of</strong> the SHH transcriptional target Gli1. The binding <strong>of</strong><br />

PN-1 to LRPs interferes with SHH-induced Cyclin D1 expression. CGNPs isolated from PN-<br />

1 deficient mice exhibit enhanced basal proliferation rates due to over-activation <strong>of</strong> the SHH<br />

pathway <strong>and</strong> show higher sensitivity to exogenous SHH. In vivo, the PN-1 deficiency alters the<br />

expression <strong>of</strong> SHH target genes. In addition, the onset <strong>of</strong> CGNP differentiation is delayed, which<br />

results in an enlarged outer External Granular Layer. Furthermore, the PN-1 deficiency leads<br />

to an overproduction <strong>of</strong> CGNPs <strong>and</strong> enlargement <strong>of</strong> the Internal Granular Layer in a subset <strong>of</strong><br />

cerebellar lobes during late development <strong>and</strong> adulthood. We propose that PN-1 contributes to<br />

shaping <strong>of</strong> the cerebellum by promoting cell cycle exit.<br />

30 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i019i<br />

Fibrin Deposition Accelerates Neurovascular Damage <strong>and</strong><br />

Neuroinflammation in Mouse Models <strong>of</strong> Alzheimer’s Disease<br />

Paul J* <strong>and</strong> Strickl<strong>and</strong> S<br />

Laboratory <strong>of</strong> Neurobiology <strong>and</strong> Genetics, The Rockefeller University, New York, New York, USA<br />

Presenting author e-mail: jpaul@rockefeller.edu<br />

Cerebrovascular dysfunction contributes to the pathology <strong>and</strong> progression <strong>of</strong> Alzheimer’s disease<br />

(AD) but the mechanisms are not completely understood. Using transgenic mouse models <strong>of</strong><br />

Alzheimer’s disease (TgCRND8 <strong>and</strong> Tg2576), we evaluated blood-brain barrier damage <strong>and</strong><br />

the role <strong>of</strong> fibrin <strong>and</strong> fibrinolysis in the progression <strong>of</strong> b-amyloid pathology. These mouse<br />

models showed age-dependent fibrin deposition coincident with areas <strong>of</strong> blood-brain barrier<br />

permeability as demonstrated by Evans blue extravasation. Three lines <strong>of</strong> evidence suggest that<br />

fibrin contributes to the pathology: 1, AD mice with only one functional plasminogen gene <strong>and</strong><br />

therefore with reduced fibrinolysis have increased neurovascular damage relative to AD mice;<br />

2 Conversely, AD mice with only one functional fibrinogen gene have decreased blood-brain<br />

barrier damage; 3, Treatment <strong>of</strong> AD mice with the plasmin inhibitor tranexamic acid aggravated<br />

pathology, while removal <strong>of</strong> fibrinogen from the circulation <strong>of</strong> AD mice with ancrod treatment<br />

attenuated measures <strong>of</strong> neuroinflammation <strong>and</strong> vascular pathology. These results suggest fibrin<br />

is a mediator <strong>of</strong> inflammation <strong>and</strong> may impede the reparative process for neurovascular damage<br />

in AD. Fibrin <strong>and</strong> the mechanisms involved in its accumulation <strong>and</strong> clearance may present novel<br />

therapeutic targets in slowing the progression <strong>of</strong> Alzheimer’s disease.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 31


i020i<br />

Anxiety-like Behavior <strong>and</strong> Impaired Fear Extinction in Mice with<br />

Altered Control <strong>of</strong> Extracellular Brain Proteolytic Activity<br />

Meins M*, Herry C, Moreno E, Fischer C, Lüthi A, Monard D<br />

Friedrich Miescher Institute for Biomedical Research, Basel, Switzerl<strong>and</strong><br />

Presenting author e-mail: mmeins@fmi.ch<br />

Regulation <strong>of</strong> serine proteases has been implicated in modulating synaptic plasticity. Protease<br />

nexin-1 (PN-1), an endogenous serine protease inhibitor, is expressed in specific neuronal<br />

subpopulations in the adult central nervous system. The in vivo expression <strong>of</strong> PN-1 is regulated<br />

by neuronal activity. In mice lacking PN-1 (PN-1 –/–), the brain homogenate proteolytic pr<strong>of</strong>ile<br />

is altered <strong>and</strong> NMDA receptor-mediated transmission is reduced. To examine the in vivo<br />

consequences <strong>of</strong> these changes, we compared cognitive <strong>and</strong> emotional responses <strong>of</strong> PN-1 –/– mice<br />

to their wildtype littermates. No memory impairments were detected. In contrast, PN-1 –/– mice<br />

displayed enhanced anxiety-like <strong>and</strong> avoidance behavior. Differences between wildtype <strong>and</strong><br />

PN-1 –/– littermates were also detected in cued fear conditioning. Reduced extinction <strong>of</strong> freezing<br />

behavior was detected in the PN-1 –/– mice. This was reflected at the cellular level by decreased<br />

immunoreactivity <strong>of</strong> c-FOS, an indicator <strong>of</strong> neuronal activity, in the PN-1 –/– basolateral<br />

amygdala. Surprisingly, increased immunoreactivity was detected in the same area in the PN-1<br />

–/– mice which underwent fear conditioning without extinction training. This may suggest that<br />

the increased reaction displayed by PN-1 –/– mice to some stressful situations may be the result<br />

<strong>of</strong> increased neuronal activation <strong>of</strong> the relevant neurons. The increased anxiety-like behavior <strong>and</strong><br />

the reduced fear extinction in the PN-1 –/– mice could be explained by an impaired modulation<br />

<strong>of</strong> NDMA receptor mediated activity in some regions <strong>of</strong> the amygdala. To investigate this<br />

hypothesis, we are analyzing downstream effectors <strong>of</strong> NMDA receptor activation in these areas.<br />

These results not only corroborate findings supporting the importance <strong>of</strong> the control <strong>of</strong> proteolytic<br />

activity in synaptic plasticity but they also suggest the PN-1 –/– mice as an interesting model to<br />

study the brain structures <strong>and</strong> the molecular mechanisms involved in fear extinction.<br />

32 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i021i<br />

Tissue <strong>Plasminogen</strong> Activator Is Co-Packaged <strong>and</strong><br />

Co-Transported to Synaptic Sites with a Key<br />

Neuromodulator Associated with Synaptic Plasticity<br />

Lochner JE* 1 , Spangler E 1 , Schuttner LC 1 , Scalettar BA 2<br />

1 Biochemistry & <strong>Molecular</strong> <strong>Biology</strong>, 2 Physics, Lewis & Clark College, Portl<strong>and</strong>, Oregon, USA<br />

Presenting author e-mail: lochner@lclark.edu<br />

Tissue plasminogen activator (tPA) is highly expressed in the hippocampus <strong>and</strong> is implicated<br />

in modifying synaptic efficacy during learning <strong>and</strong> memory. tPA is postulated to exert its<br />

influence on synaptic plasticity by initiating an extracellular proteolytic cascade that promotes the<br />

conversion <strong>of</strong> precursor brain-derived neurotrophic factor (proBDNF) to mature BDNF (mBDNF)<br />

extracellularly, at synapses. mBDNF is a potent neuromodulator that is known to augment<br />

synapse density <strong>and</strong> elicit long-lasting enhancement <strong>of</strong> synaptic transmission. Motivated by<br />

recent interest in possible interactions between tPA <strong>and</strong> proBDNF, we have developed a system<br />

that facilitates study <strong>of</strong> the packaging, transport, <strong>and</strong> secretion <strong>of</strong> these two neuromodulatory<br />

proteins in cultured hippocampal neurons. We find 89% colocalization <strong>of</strong> tPA <strong>and</strong> proBDNF<br />

chimeras within dense-core granules (DCGs) <strong>of</strong> mature hippocampal neurons, indicating efficient<br />

co-packaging <strong>of</strong> these neuromodulators. Additionally, we find that these two neuromodulators<br />

undergo rapid co-transport (at speeds <strong>of</strong> ~1 µm/s) along neuronal processes, suggesting that tPA<br />

<strong>and</strong> pro-BDNF are co-delivered to synaptic sites in DCGs. Evaluation <strong>of</strong> the colocalization <strong>of</strong> these<br />

two neuromodulators at the organelle level in post-synaptic dendritic spines reveals that tPA<br />

<strong>and</strong> proBDNF chimeras colocalize in individual DCGs in 27% <strong>of</strong> spines, <strong>and</strong> that 80% <strong>of</strong> spines<br />

which contain DCGs contain both chimeras in the same DCG. Our results suggest that efficient<br />

co-packaging <strong>and</strong> co-transport produces a pool <strong>of</strong> DCGs containing both tPA <strong>and</strong> proBDNF at<br />

post-synaptic sites, where these neuromodulators can undergo activity-dependent co-release from<br />

DCGs <strong>and</strong> then interact <strong>and</strong>/or mediate changes that influence learning <strong>and</strong> memory.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 33


i022i<br />

Tissue <strong>Plasminogen</strong> Activator Modulates <strong>Cellular</strong><br />

<strong>and</strong> Behavioral Response to Cocaine<br />

Maiya R* 1 , Zhou Y 2 , Norris EH 1 , Kreek MJ 2 , Strickl<strong>and</strong> S 1<br />

1 2 Laboratory <strong>of</strong> Neurobiology <strong>and</strong> Genetics, Laboratory <strong>of</strong> the <strong>Biology</strong> <strong>of</strong> Addictive Diseases, The<br />

Rockefeller University, New York, New York, USA<br />

*Presenting author e-mail: rmaiya@mail.rockefeller.edu<br />

Cocaine exposure induces long lasting molecular <strong>and</strong> cellular adaptations in the brain. Tissue<br />

plasminogen activator (tPA), an extracellular protease implicated in regulating neuronal plasticity,<br />

may orchestrate some <strong>of</strong> these cocaine-induced neuro-adaptations. In this study, the effects<br />

<strong>of</strong> acute <strong>and</strong> chronic cocaine administration on tPA activity in the brain were examined. Mice<br />

were injected with cocaine in the acute (3 injections <strong>of</strong> 15 mg/kg) or chronic (3x15mg/kg for 14<br />

days) ‘’binge’’ paradigm. In situ <strong>and</strong> in-gel zymography revealed that in the amygdala <strong>of</strong> WT<br />

animals, acute cocaine exposure increased tPA activity whereas chronic cocaine administration<br />

decreased tPA activity. Acute cocaine also elevated levels <strong>of</strong> plasma corticosterone, corticotropin<br />

releasing factor (CRF), <strong>and</strong> CRF receptor 1 (CRF-R1) mRNAs in the amygdala <strong>of</strong> both WT <strong>and</strong><br />

tPA–/– animals. Phosphorylation <strong>of</strong> extracellular signal regulated kinase 1/2 (ERK1/2), a marker<br />

for post-synaptic plasticity events, was observed in the amygdala <strong>of</strong> WT but not tPA–/– mice<br />

after acute cocaine exposure. In comparison to WT animals, tPA–/– animals showed attenuated<br />

c-fos expression in the amygdala after acute cocaine administration. These results suggest altered<br />

cocaine-induced neuro-adaptation in the amygdala <strong>of</strong> tPA–/– mice. Cocaine-induced anxiety<br />

was also examined in both WT <strong>and</strong> tPA–/– animals using the elevated plus maze. In contrast to<br />

WT animals, tPA–/– animals were resistant to the anxiogenic effects <strong>of</strong> acute cocaine. After acute<br />

cocaine, tPA–/– mice displayed increased total arm entries in comparison to WT mice, suggesting<br />

enhanced cocaine-induced locomotor stimulation. In summary, these results suggest a significant<br />

role for tPA in the modulation <strong>of</strong> cocaine-induced behaviors <strong>and</strong> neuronal plasticity.<br />

34 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i023i<br />

Characterisation <strong>of</strong> the Pathway <strong>of</strong> Polymerisation<br />

<strong>of</strong> Wildtype Neuroserpin <strong>and</strong> the Ser49Pro Mutant<br />

that Underlies the Dementia FENIB<br />

Hägglöf P*, Belorgey D, Karlsson-Li S, Sharp LK, Lomas DA<br />

Department <strong>of</strong> Medicine, University <strong>of</strong> Cambridge, Cambridge Institute for Medical Research,<br />

Cambridge, United Kingdom<br />

Presenting author e-mail: pmh43@cam.ac.uk<br />

Neuroserpin is expressed during the late stage <strong>of</strong> development in neurons <strong>of</strong> the central <strong>and</strong><br />

peripheral nervous system <strong>and</strong> in the adult brain. The target proteinase <strong>of</strong> neuroserpin is tissue<br />

plasminogen activator (tPA) <strong>and</strong> it is likely to be important in the control <strong>of</strong> synaptic plasticity,<br />

in learning, memory <strong>and</strong> can act as a neuroprotectant. It can decrease the toxicity <strong>of</strong> the Ab1-<br />

42 peptide that is central to the pathogenesis <strong>of</strong> Alzheimer’s disease. Point mutations in the<br />

neuroserpin gene underlie the autosomal dominant dementia, familial encephalopathy with<br />

neuroserpin inclusion bodies, (FENIB). This is characterised by the retention <strong>of</strong> ordered polymers<br />

<strong>of</strong> neuroserpin within the endoplasmic reticulum <strong>of</strong> neurons. We have previously shown that<br />

polymers result from the sequential linkage between the reactive centre loop <strong>of</strong> one neuroserpin<br />

molecule <strong>and</strong> b-sheet A <strong>of</strong> another. We show here that the polymerisation <strong>of</strong> both wildtype <strong>and</strong><br />

the Ser49Pro mutant <strong>of</strong> neuroserpin that causes FENIB is dependent <strong>of</strong> pH demonstrating that<br />

the His-338 residue in the shutter region is important for in the opening <strong>of</strong> b-sheet A. Tryptophan,<br />

extrinsic fluorescence <strong>and</strong> fluorphore labelled cysteine mutants suggest that the RCL may be<br />

preinserted in b-sheet A <strong>and</strong> then expelled during in an at least two steps in the formation <strong>of</strong><br />

polymers: a fast unimolecular process that is likely to reflect opening <strong>of</strong> b-sheet A followed by<br />

the slower process <strong>of</strong> polymer formation. Taken together these data suggest a two-step kinetic<br />

mechanism for the formation <strong>of</strong> the polymers <strong>of</strong> neuroserpin that underlie the dementia FENIB.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 35


i024i<br />

Matriptase Is an Essential Inhibitory Target for<br />

Hepatocyte Growth Factor Activator Inhibitor-1 during<br />

both Embryonic Development <strong>and</strong> Postnatal Life<br />

Szabo R*, Molinolo A, List K, Bugge TH<br />

Oral <strong>and</strong> Pharyngeal Cancer Branch, NIDCR, NIH, Bethesda, Maryl<strong>and</strong>, USA<br />

Presenting author e-mail: rszabo@nidcr.nih.gov<br />

Hepatocyte growth factor activator inhibitor-1 (HAI-1) is a Kunitz-type transmembrane serine<br />

protease inhibitor proposed to inhibit the activity <strong>of</strong> several trypsin-like serine proteases,<br />

including hepsin, hepatocyte growth factor activator, prostasin (CAP1/PRSS8), <strong>and</strong> matriptase.<br />

Here we generated HAI-1-deficient mice to determine the biological function <strong>of</strong> HAI-1 <strong>and</strong><br />

identify physiological HAI-1 inhibitory targets during development <strong>and</strong> postnatal life. HAI-<br />

1-deficient mice died at mid-gestation due to placental insufficiency caused by a disruption <strong>of</strong><br />

the epithelial integrity <strong>of</strong> placental chorionic trophoblasts associated with chorionic basement<br />

membrane dissolution, loss <strong>of</strong> E-cadherin, <strong>and</strong> loss <strong>of</strong> membrane-associated b-catenin.<br />

Interestingly, however, matriptase gene ablation in HAI-1-deficient embryos restored the integrity<br />

<strong>of</strong> chorionic trophoblasts <strong>and</strong> enabled both normal placentation <strong>and</strong> development to term. HAI-<br />

1 was recently reported to also be required for mouse postnatal survival beyond two weeks. To<br />

determine the inhibitory targets for HAI-1 during postnatal life, we generated HAI-1 null mice<br />

in a matriptase hypomorphic background, which display 82 - >99 % reduction in matriptase<br />

mRNA levels in epithelial tissues. Low matriptase levels enabled both embryonic development<br />

<strong>and</strong> normal postnatal survival <strong>of</strong> HAI-1-deficient mice when followed for up to six months. Taken<br />

together, this study identifies matriptase as an essential inhibitory target for HAI-1 during both<br />

embryonic development <strong>and</strong> postnatal life.<br />

36 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i025i<br />

Mice with very low Matriptase Are Viable <strong>and</strong> Phenocopy<br />

Human Autosomal Ichthyosis with Hypotrichosis Syndrome<br />

List K* 1 , Currie B 1 , Scharschmidt T 2 , Szabo R 1 , Molinolo A 1 , Shireman J 1 , Segre J 2 , Bugge TH 1<br />

1 Oral <strong>and</strong> Pharyngeal Cancer Branch, NIDCR, NIH, Bethesda, Maryl<strong>and</strong>, USA;<br />

2 National Human Genome Research Institute, NIH, Bethesda, Maryl<strong>and</strong>, USA<br />

Presenting author e-mail: klist@nidcr.nih.gov<br />

Complete deficiency <strong>of</strong> the type II transmembrane serine protease matriptase or its c<strong>and</strong>idate<br />

downstream target, the GPI-anchored serine protease prostasin, causes epidermal barrier<br />

disruption <strong>and</strong> neonatal death <strong>of</strong> mice. Here we report the surprising observation that mice with a<br />

one-hundred-fold reduction in epidermal matriptase mRNA levels, generated by inserting a leaky<br />

splice acceptor site into intron 1 <strong>of</strong> the matriptase gene, are viable <strong>and</strong> fertile. Interestingly, these<br />

matriptase hypomorphic mice display skin <strong>and</strong> hair defects that are identical to those <strong>of</strong> humans<br />

with Autosomal Recessive Ichthyosis with Hypotrichosis (ARIH): an inherited disorder recently<br />

linked to homozygosity for a G827R substitution in the matriptase serine protease domain. Thus,<br />

matriptase hypomorphic mice developed marked hyperkeratosis with impaired desquamation<br />

<strong>and</strong> focal dermal inflammation, as well as hypotrichosis with brittle, thin, uneven, <strong>and</strong> sparse<br />

hair. Proteomic analysis <strong>of</strong> matriptase hypomorphic epidermis revealed markedly reduced<br />

prostasin activation <strong>and</strong> pr<strong>of</strong>ilaggrin proteolytic processing. This work strongly supports reduced<br />

matriptase activity as one etiological origin <strong>of</strong> human ARIH <strong>and</strong> provides a mouse model for the<br />

exploration <strong>of</strong> matriptase in other physiological <strong>and</strong> pathological processes, including epithelial<br />

carcinogenesis.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 37


i026i<br />

Role <strong>of</strong> Urokinase-Receptor in Hematopoietic Stem Cell Trafficking<br />

Montuori N 1 , Selleri C 2 , Ricci P 2 , Visconte V 1 , Carriero MV 3 , Rotoli B 2 , Rossi G 1 , Ragno P* 4<br />

1 Department <strong>of</strong> <strong>Cellular</strong> <strong>and</strong> <strong>Molecular</strong> <strong>Biology</strong> <strong>and</strong> Pathology, 2 Department <strong>of</strong> Biochemistry <strong>and</strong><br />

Medical Biotechnology, ‘’Federico II’’ University, Naples, Italy;<br />

3 Department <strong>of</strong> Experimental Oncology, National Cancer Institute, Naples, Italy;<br />

4 Department <strong>of</strong> Chemistry, University <strong>of</strong> Salerno, Salerno, Italy<br />

Presenting author e-mail: pragno@unisa.it<br />

Cleaved forms <strong>of</strong> soluble urokinase-receptor (c-suPAR) have been detected in body fluids from<br />

patients affected by various tumors. We reported increased c-suPAR levels in sera <strong>of</strong> healthy<br />

donors during granulocyte colony-stimulating factor (G-CSF)-induced mobilization <strong>of</strong> CD34+<br />

hematopoietic stem cells (HSCs). In vitro, c-suPAR or its derived chemotactic peptide (uPAR84-<br />

95) stimulated migration <strong>of</strong> human CD34+ HSCs <strong>and</strong> inactivated CXCR4, the chemokine receptor<br />

primarily responsible for HSC retention in bone marrow (BM). These results suggested that<br />

c-suPAR could potentially contribute to regulate HSC trafficking from <strong>and</strong> to BM. Therefore, we<br />

investigated uPAR84-95 effects on mobilization <strong>of</strong> mouse CD34+ hematopoietic stem/progenitor<br />

cells (HSCs/HPCs). We first demonstrated that uPAR84-95 stimulated in vitro dose-dependent<br />

migration <strong>of</strong> mouse CD34+ M1 leukemia cells <strong>and</strong> inactivated murine CXCR4. uPAR84-95<br />

capability to induce mouse HSC/HPC release from bone marrow <strong>and</strong> migration into the<br />

circulation was then investigated in vivo. uPAR84-95 intraperitoneal administration induced rapid<br />

leukocytosis, which was associated with an increase in peripheral blood CD34+ HSCs/HPCs. In<br />

vitro colony assays confirmed that uPAR84-95 mobilized hematopoietic progenitors, showing an<br />

increase in circulating colony-forming cells.<br />

We are now studying suPAR <strong>and</strong> c-suPAR effects on HSCs, to investigate their role in HSC<br />

retention in BM, <strong>and</strong>, therefore, their potential contribution also to the homing <strong>and</strong> the<br />

engraftment <strong>of</strong> HSCs to the BM. In vitro effects <strong>of</strong> anti-uPAR antibodies <strong>and</strong> <strong>of</strong> suPAR <strong>and</strong> csuPAR<br />

on adhesion <strong>and</strong> proliferation <strong>of</strong> human G-CSF-mobilized <strong>and</strong> BM HSCs have been<br />

examined in long-term cultures (LTC), which represent the best surrogate in vitro <strong>of</strong> BM<br />

hematopoiesis.<br />

38 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i027i<br />

<strong>Plasminogen</strong> Activator Inhibitor-1, PAI-1, Regulates<br />

the Akt Survival Pathway<br />

Rømer MU, Larsen L, Offenberg H, Brünner N, Lademann U*<br />

Department <strong>of</strong> Veterinary Pathobiology, Faculty <strong>of</strong> Life Science, University <strong>of</strong> Copenhagen, Frederiksberg,<br />

Denmark<br />

Presenting auther e-mail: ul@life.ku.dk<br />

<strong>Plasminogen</strong> Activator Inhibitor-1, PAI-1, inhibits the activation <strong>of</strong> the plasminogen activation<br />

system, which is involved in cancer growth <strong>and</strong> dissemination. Increased tumor tissue levels <strong>of</strong><br />

PAI-1 would therefore be expected to inhibit cancer cell progression. However, PAI-1 is elevated<br />

in many solid tumors <strong>and</strong> this elevation has consistently been shown to be associated with shorter<br />

patient survival. The reason for this has remained unclear.<br />

The phosphatidylinositol-3 kinase (PI3K)/Akt pathway is overactivated in a wide range <strong>of</strong> tumor<br />

types, <strong>and</strong> this triggers a cascade <strong>of</strong> responses, including inhibition <strong>of</strong> apoptosis. PAI-1 has been<br />

reported to modulate the PI3 kinase /Akt pathway, thus it is possible that PAI-1 inhibits apoptosis<br />

<strong>of</strong> tumor cells via activation <strong>of</strong> this pathway.<br />

We have previously shown that PAI-1–/– fibrosarcoma cells are significantly more sensitive than<br />

wild-type fibrosarcoma cells to etoposide-induced apoptosis. To further study the signalling<br />

pathways involved in the anti-apoptotic function <strong>of</strong> PAI-1 we asked the question whether Akt<br />

activation is involved in PAI-1 mediated inhibition <strong>of</strong> apoptosis. We demonstrate that Akt is<br />

hyperfosforylated in wild type fibrosarcoma cells (endogenous PAI-1 expression) compared<br />

with PAI-1–/– fibrosarcoma cells. This hyperphosphorylation is accompanied by increased<br />

phosphorylation <strong>of</strong> downstream targets <strong>of</strong> Akt. When wild type fibrosarcoma cells were treated<br />

with a specific inhibitor against Akt1/2 (Akt inhibitor VIII) <strong>and</strong> PI3K (LY294002) it induced a<br />

sensitizing effect on etoposide induced cell death, further supporting an apoptosis regulating role<br />

<strong>of</strong> the PI3K/Akt survival pathway in our cell model.<br />

Altogether, our data suggests that PAI-1 inhibits etoposide induced cell death via activation <strong>of</strong> the<br />

Akt survival pathway.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 39


i028i<br />

A Host <strong>Plasminogen</strong> Activator Inhibitor-1 Deficiency Promotes<br />

Proliferation <strong>and</strong> Resistance to Apoptosis by <strong>Activation</strong><br />

<strong>of</strong> the PI3-K/Akt Pathway in Endothelial Cells<br />

Balsara RD* 1,2,3 , Castellino FJ 1,2,3 , Ploplis VA 1,2,3<br />

1 2 3 W. M. Keck Center for Transgene Research, Department <strong>of</strong> Chemistry <strong>and</strong> Biochemistry, Notre Dame<br />

Cancer Institute, University <strong>of</strong> Notre Dame, Indiana, USA<br />

Presenting author e-mail: rbalsara@nd.edu<br />

Angiogenic function <strong>of</strong> plasminogen activator inhibitor-1 (PAI-1) <strong>and</strong> its underlying mechanism<br />

were studied utilizing endothelial cells (EC) isolated from the aortas <strong>of</strong> wild-type (WT) <strong>and</strong> PAI-<br />

1-deficient mice. The enhanced proliferation <strong>of</strong> PAI-1–/– EC was associated with hyperactivation<br />

<strong>of</strong> Akt(P-Ser473) <strong>and</strong> accompanied by resistance to apoptosis. Disruption <strong>of</strong> VEGF/VEGFR-1<br />

interaction attenuated activation <strong>of</strong> Akt(P-Ser473) in PAI-1–/– EC. Involvement <strong>of</strong> the PI3k/Akt<br />

survival axis in both WT <strong>and</strong> PAI-1–/– EC was discerned by treating cells with the<br />

pharmacological inhibitor wortmannin, which also diminished cell proliferation. Furthermore,<br />

the higher levels <strong>of</strong> Akt(P-Ser473) in PAI-1–/– EC were regulated by higher levels <strong>of</strong> inactive<br />

phosphatase PTEN. In PAI-1–/– EC, increased levels <strong>of</strong> Akt(P-Ser473) were associated with higher<br />

levels <strong>of</strong> inactive caspase-9, which is a direct phosphorylation target <strong>of</strong> Akt. Subsequently lower<br />

protein levels <strong>and</strong> activity <strong>of</strong> pro- <strong>and</strong> active-caspase-3 were detected in PAI-1–/– EC. Treatment<br />

<strong>of</strong> PAI-1–/– EC with exogenous r-PAI-1 controlled the increased rate <strong>of</strong> cell proliferation with<br />

concomitant modulation <strong>of</strong> the components <strong>of</strong> the Akt pathway. In the presence <strong>of</strong> r-PAI-1, levels<br />

<strong>of</strong> inactive PTEN(P-Ser380) <strong>and</strong> Akt(P-Ser473) were decreased, <strong>and</strong> the levels <strong>of</strong> pro- <strong>and</strong> activecaspase-3<br />

were increased. A r-PAI-1 mutant that had compromised binding ability to low density<br />

lipoprotein receptor-related protein (LRP) failed to alter cell proliferation <strong>of</strong> PAI-1–/– cells. These<br />

results indicate that PAI-1 mediates its anti-proliferative activity in an LRP-dependent manner.<br />

Additionally, PAI-1 plays a pivotal role in modulating the Akt pathway, which is an important<br />

participant in transducing signaling events during angiogenesis.<br />

40 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i029i<br />

Bomapin Is a Redox-regulated Serpin which Stabilizes Retinoblastoma<br />

Protein during Apoptosis <strong>and</strong> Increases Proliferation <strong>of</strong> Leukemia Cells<br />

Przygodzka P, Olausson B, Tengel T, Larsson G, Wilczynska M*<br />

Department <strong>of</strong> Medical Biochemistry <strong>and</strong> Biophysics, Umeå University, Umeå, Sweden<br />

Presenting author e-mail: Malgorzata.Wilczynska@medchem.umu.se<br />

Bomapin is a hematopoietic- <strong>and</strong> leukemia-specific intracellular serine protease inhibitor.<br />

Its expression is restricted to bone marrow <strong>and</strong> to peripheral blood leukocytes <strong>of</strong> patients<br />

with myeloid leukemias. Consistent with this, bomapin is also constitutively expressed in<br />

promyelocytic <strong>and</strong> monocytic cell lines (HL60, THP1, <strong>and</strong> AML-193). Herein we show that<br />

naturally expressed bomapin is localized in the nucleus, where it is stabilized by a disulfide<br />

bond linking the only two bomapin cysteines. Computer modeling has shown that the cysteines<br />

are distant in the reduced bomapin, but can easily be disulfide-linked without distortion <strong>of</strong><br />

the overall bomapin structure. Stable, ectopic expression <strong>of</strong> wild-type bomapin in K562 cells<br />

increased cell proliferation by about 50%. Importantly, the enhancement <strong>of</strong> cell proliferation was<br />

not observed for a bomapin mutant lacking the disulfide bond. We show also that in proliferating<br />

leukemia cells, bomapin existed in a multi-protein complex including lamin-A/C, retinoblastoma<br />

(Rb) protein, beta-actin, <strong>and</strong> non-muscle myosin IIA. Bomapin expression resulted in stabilization<br />

<strong>of</strong> Rb protein during serum deprivation-induced apoptosis. We propose that bomapin is a<br />

redox-regulated serpin which enhances proliferation potential <strong>of</strong> leukemia cells via assembly<br />

into a multi-protein complex including also the Rb protein, <strong>and</strong> stabilizes Rb protein against<br />

degradation during apoptosis.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 41


i030i<br />

Hespin as a Cell Survival Factor<br />

Qiu D, Owen K, Edwards DR, Ellis V*<br />

Biomedical Research Centre, School <strong>of</strong> Biological Sciences, University <strong>of</strong> East Anglia, Norwich, United<br />

Kingdom<br />

Presenting author e-mail: v.ellis@uea.ac.uk<br />

Hepsin is a type II transmembrane serine protease upregulated in a variety <strong>of</strong> cancers <strong>and</strong> may<br />

be functionally involved in disease progression. How hepsin might mediate these effects <strong>and</strong><br />

via which substrates is currently unknown. Our qRT-PCR data show that hepsin expression<br />

in prostate cancer tissue is positively correlated with both malignancy (p


i031i<br />

Urokinase (uPA) Protects Endothelial Cell against Apoptosis by<br />

Upregulating the X-linked Inhibitor <strong>of</strong> Apoptosis Protein (XIAP)<br />

Prager GW* 1,2 , Koschelnick Y 1 , Mihaly J 1 , Brunner P 1 , Binder BR 1<br />

1 Department <strong>of</strong> Vascular <strong>Biology</strong> <strong>and</strong> Thrombosis Research, Centre for Bio-<strong>Molecular</strong> Medicine <strong>and</strong><br />

Pharmacology, 2 Clinical Division <strong>of</strong> Oncology, Department <strong>of</strong> Medicine I <strong>and</strong> Cancer Center, Medical<br />

University Vienna, Austria<br />

Presenting author e-mail: gerald.prager@meduniwien.ac.at<br />

uPA has originally been implicated to assist the angiogenic process by it’s proteolytic properties.<br />

It is now becoming increasingly evident that uPA additionally elicits many pro-angiogenic<br />

responses like differentiation, proliferation <strong>and</strong> cell migration in a non-proteolytic fashion via<br />

induction <strong>of</strong> intracellular signal transduction.<br />

We now demonstrate that in endothelial cells uPA protects against apoptosis by transcriptional<br />

upregulation <strong>of</strong> inhibitor <strong>of</strong> apoptosis proteins (IAPs), among them most prominently the Xlinked<br />

inhibitor <strong>of</strong> apoptosis protein (XIAP). In contrast to canonical growth factors, like vascular<br />

endothelial growth factor (VEGF), uPA elicits anti-apoptosis independently <strong>of</strong> the PI3-kinase<br />

pathway. uPA-induced cell survival is dependent on the type <strong>of</strong> extracellular matrix used<br />

indicating the involvement <strong>of</strong> integrin adhesion receptors. Thereby, uPA induces phosphorylation<br />

<strong>of</strong> the CDC42 downstream effector p21-activated kinase 1 (PAK1) <strong>and</strong> phosphorylation <strong>of</strong><br />

IkappaB kinase alpha (IKKa), which itself induces NFkappaB activation. Blocking NFkappaB by<br />

the use <strong>of</strong> the specific NFkappaB inhibitor BAY 11-7082 or by adenoviral-mediated overexpression<br />

<strong>of</strong> its inhibitor IkB, inhibits uPA-induced XIAP expression as well as uPA-induced cell survival.<br />

Furthermore, silencing XIAP expression by siRNA significantly reduces cell survival efficiencies<br />

<strong>of</strong> uPA in endothelial cells.<br />

From these data we conclude that uPA, which is a main player in endothelial cell migration <strong>and</strong><br />

invasion, provides an additional, PI3-kinase independent autocrine or paracrine cell survival<br />

mechanism.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 43


i032i<br />

Pro-fibrinolytic Effects <strong>of</strong> Metalloproteinases during Skin<br />

Wound Healing in the Absence <strong>of</strong> <strong>Plasminogen</strong><br />

Lund LR* 1 , Green KA 1 , Almholt K 1 , Ploug M 1 , Bugge TH 2 , Rømer J 1<br />

1 Finsen Laboratory, Rigshospitalet, Copenhagen Biocenter, Copenhagen, Denmark;<br />

2Proteases <strong>and</strong> Tissue Remodeling Unit, National Institute <strong>of</strong> Dental <strong>and</strong> Crani<strong>of</strong>acial Research, National<br />

Institutes <strong>of</strong> Health, Bethesda, Maryl<strong>and</strong>, USA<br />

Presenting author e-mail: lund@inet.uni2.dk<br />

Genetic ablation <strong>of</strong> plasminogen as well as pharmacological inhibition <strong>of</strong> metalloproteinase<br />

activity delays skin wound healing in mice, while the combined inhibition <strong>of</strong> these two<br />

enzyme systems completely prevents healing. In the present study the impact <strong>of</strong> plasmin <strong>and</strong><br />

metalloproteinases as pro-fibrinolytic enzymes has been investigated by comparing skin wound<br />

healing in the absence <strong>and</strong> presence <strong>of</strong> fibrin. <strong>Plasminogen</strong>-deficiency impairs skin wound<br />

healing kinetics, but this delay is only partially restored in the absence <strong>of</strong> fibrin. This suggests that<br />

plasmin-mediated fibrinolysis is the primary, but not the exclusive, requirement for healing <strong>of</strong><br />

wounds in these mice. In addition, we observe that lack <strong>of</strong> fibrin reduces plasminogen activation<br />

significantly during wound healing. The pro-fibrinolytic role <strong>of</strong> metalloproteinases is revealed<br />

by the finding that lack <strong>of</strong> fibrin partially restores the otherwise arrested healing <strong>of</strong> plasminogendeficient<br />

wounds after metalloproteinase-inhibition. In conclusion, the residual impairment<br />

<strong>of</strong> skin wound healing in the absence <strong>of</strong> fibrin suggests the existence <strong>of</strong> a fibrin-independent<br />

substrate(s) for plasmin <strong>and</strong> metalloproteinases. Furthermore, these in vivo data reveal that<br />

galardin-sensitive metalloproteinases mediate compensatory fibrinolysis to facilitate wound<br />

healing in the absence <strong>of</strong> plasmin.<br />

44 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i033i<br />

Complementary Roles <strong>of</strong> Intracellular <strong>and</strong> Pericellular Collagen<br />

Degradation Pathways in Mesenchymal Cell Survival <strong>and</strong> Proliferation<br />

Wagenaar-Miller RA 1 , Engelholm LH 2 , Gavard J 1 , Yamada S 3 , Gutkind JS 1 , Behrendt N 2 , Holmbeck K 3 ,<br />

Bugge TH* 1<br />

1 Oral <strong>and</strong> Pharyngeal Cancer Branch, NIDCR, NIH, Bethesda, Maryl<strong>and</strong>, USA;<br />

2 Finsen Laboratory, Copenhagen, Denmark;<br />

3 Skeletal Morphogenesis Branch, NIDCR, NIH, Bethesda, Maryl<strong>and</strong>, USA<br />

Presenting author e-mail: thomas.bugge@nih.gov<br />

Mesenchymal cells must be capable <strong>of</strong> proteolytically remodeling their collagen-rich pericellular<br />

environments to survive, proliferate, <strong>and</strong> properly differentiate. Two key turnover pathways<br />

have been described for collagen: intracellular cathepsin-mediated degradation <strong>and</strong> pericellular<br />

collagenase-mediated degradation. Here we show that intracellular <strong>and</strong> pericellular collagen<br />

turnover pathways have complementary functions in collagen remodeling. Combined, but<br />

not individual, deficits in intracellular collagen degradation (uPARAP/Endo180 ablation) <strong>and</strong><br />

pericellular collagen degradation (MT1-MMP ablation) caused proliferative failure <strong>and</strong> poor<br />

survival <strong>of</strong> osteoblasts <strong>and</strong> chondrocytes within their respective collagen type I- <strong>and</strong> collagen<br />

type II-rich osteogenic <strong>and</strong> chondrogenic niches. This severely impaired overall bone formation,<br />

leading to uniform postnatal death <strong>of</strong> combined uPARAP/Endo180;MT1-MMP-deficient<br />

mice. These findings have important implications for the use <strong>of</strong> pharmacological inhibitors <strong>of</strong><br />

collagenase activity aimed at preventing connective tissue destruction in a variety <strong>of</strong> diseases.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 45


i034i<br />

Interplay between MMPs <strong>and</strong> the Endocytic Collagen<br />

Receptor, uPARAP/Endo180, in Collagen Degradation<br />

Behrendt N* 1 , Madsen DH 1 , Ingvarsen S 1 , Hillig T 1 , Wagenaar-Miller R 2 , Kjøller L 1 , Gårdsvoll H 1 ,<br />

Høyer-Hansen G 1 , Bugge TH 2 , Engelholm LH 1<br />

1 The Finsen Laboratory, Rigshospitalet, Copenhagen, Denmark;<br />

2Oral & Pharyngeal Cancer Branch, National Institute <strong>of</strong> Dental <strong>and</strong> Crani<strong>of</strong>acial Research, NIH,<br />

Bethesda, Maryl<strong>and</strong>, USA<br />

Presenting author e-mail: niels.behrendt@finsenlab.dk<br />

The uPAR-associated protein (uPARAP/Endo180) is an endocytic cell surface receptor on<br />

mesenchymal cells that is centrally engaged in the turnover <strong>of</strong> collagens <strong>of</strong> several subtypes. It<br />

has recently become clear that uPARAP/Endo180-mediated collagen degradation takes place<br />

both during bone development <strong>and</strong> in the invasive growth <strong>of</strong> malignant tumors. In this work, we<br />

have focused on the interplay between this new mechanism <strong>and</strong> the more extensively studied,<br />

extracellular routes <strong>of</strong> collagenolysis. In collagen internalization assays, collagen that has been<br />

pre-cleaved by a mammalian collagenase is taken up much more efficiently than intact, native<br />

collagen by uPARAP/Endo180 positive cells. This preference is governed by the acquisition <strong>of</strong><br />

a gelatin like structure <strong>of</strong> collagen, occurring upon collagenase mediated cleavage under native<br />

conditions. Furthermore, the growth <strong>of</strong> uPARAP/Endo180 deficient fibroblasts on a native<br />

collagen matrix leads to a dramatic accumulation <strong>of</strong> large collagen fragments in the culture<br />

supernatant. In contrast, wildtype fibroblasts possess the ability to direct a complete collagen<br />

breakdown sequence, including both initial, extracellular cleavage, endocytic (uPARAP/Endo180<br />

mediated) uptake <strong>of</strong> large, defined fragments <strong>and</strong> a final, intracellular degradation step. Thus,<br />

our work shows that extracellular collagenolysis <strong>and</strong> endocytic collagen turnover can occur as<br />

an integrated mechanism in cultured fibroblasts <strong>and</strong> fibroblast-like cells. Most likely, a similar<br />

mechanism is operative in the stromal cell types that are responsible for collagenolysis during<br />

tumor invasion.<br />

46 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i035i<br />

The Urokinase Receptor Ko Mice Have Reduced Keratinocytes<br />

Proliferation <strong>and</strong> Migration during Wound Healing <strong>and</strong><br />

Are Protected in a Skin Carcinogenesis Protocol<br />

D’Alessio S, Mazzieri R, Gerasi L, Blasi F*<br />

H. San Raffaele <strong>and</strong> IFOM (FIRC Institute <strong>of</strong> <strong>Molecular</strong> Oncology), Milan, Italy<br />

Presenting author e-mail: Blasi.Francesco@hsr.it<br />

We have found that the urokinase receptor (uPAR) is implicated in multistage mouse skin<br />

carcinogenesis, since uPAR Ko mice show a reduced susceptibility to tumor formation after a<br />

DMBA-TPA protocol, compared to wt-mice. The uPAR Ko mice have 70% less <strong>and</strong> much smaller<br />

papillomas. However, the rate <strong>of</strong> conversion <strong>of</strong> papillomas to carcinomas, is not different between<br />

wt <strong>and</strong> Ko mice, indicating that the role <strong>of</strong> uPAR is in the first stage <strong>of</strong> tumorigenesis. We also<br />

have noticed that uPAR is also involved in the regulation <strong>of</strong> wound healing, as uPAR Ko mice<br />

display a slight delay in wound closure compared to wt-skin. In order to directly underst<strong>and</strong><br />

the role <strong>of</strong> endogenous uPAR in these two in vivo events, we examined uPAR null-<strong>and</strong> wildtype<br />

keratinocytes behaviour in vitro. Under non–permissive conditions uPAR wt grow faster<br />

than uPAR Ko keratinocytes (<strong>and</strong> differentiate faster). After a strong growth stimulation, such<br />

as addition <strong>of</strong> EGF or serum, the difference becomes even larger as the uPAR Ko keratinocytes<br />

do not respond to EGF. Moreover EGF induces cell growth <strong>and</strong> activates ERK/MAP kinase<br />

in murine keratinocytes only when these cells express uPAR even though the EGF receptor<br />

(EGFR) expression level was not different uPAR Ko <strong>and</strong> wt cells, both in the absence <strong>and</strong> in the<br />

presence <strong>of</strong> EGF. Moreover EGFR tyrosine phosphorylation, which is involved in cell growth, was<br />

specifically decreased in uPAR Ko expressing cells both in the absence <strong>and</strong> in the presence <strong>of</strong> EGF.<br />

In terms <strong>of</strong> cell proliferation, the rescue <strong>of</strong> the uPAR +/+ phenotype was achieved by expressing<br />

murine, but not human, uPAR cDNA, in uPAR Ko cells. This suggests a role for also for uPA in<br />

EGFR activation, possibly through an autocrine signalling, as murine uPA cannot bind the human<br />

uPAR.<br />

Besides a reduced growth rate in vitro, uPAR-deficient keratinocytes are also unable to migrate,<br />

adhere <strong>and</strong> spread on a glass substrate because they fail to produce <strong>and</strong> secrete their own LN-5<br />

substrate, which is an important requirement for reepithelialization <strong>of</strong> cutaneous wounds. In fact,<br />

when we generated full-thickness excision wounds in the skin, immunological analysis <strong>of</strong> frozen<br />

sections 3 days after injury, revealed that LN5 was highly expressed in uPAR wt but was almost<br />

absent in uPAR-deficient mice, meaning that also in vivo LN5 deposition is controlled by the uPAR<br />

genotype. This effect can be observed at both RNA <strong>and</strong> protein level.<br />

In summary, our data provide evidence that uPAR controls cell migration <strong>and</strong> proliferation both<br />

in vivo <strong>and</strong> in vitro under stressed conditions such as tumor growth <strong>and</strong> wound healing.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 47


i036i<br />

Distinct Roles <strong>of</strong> Plasmin in Staphylococcus aureus-induced<br />

Sepsis <strong>and</strong> Infection Models<br />

Guo Y*, Li J, Hagström E, Ny T<br />

Department <strong>of</strong> Medical Biochemistry <strong>and</strong> Biophysics, Umeå University, Umeå, Sweden<br />

Presenting author e-mail: yong-zhi.guo@medchem.umu.se<br />

Plasmin is an important mediator in inflammatory cell migration <strong>and</strong> activation during host<br />

defense against bacterial infection. However, the exact functional role <strong>of</strong> plasmin/ogen during<br />

infection <strong>and</strong> sepsis is unclear. We have investigated the role <strong>of</strong> plasmin in murine Staphylococcus<br />

aureus-induced sepsis <strong>and</strong> infection models by using wild-type (plg +/+ ) <strong>and</strong> plasminogen-deficient<br />

(plg –/– ) mice. Our results showed that when mice were injected intravenously with 1 × 10 7 CFU <strong>of</strong><br />

S. aureus to induce systemic infection, the plg +/+ control mice had a 21-day survival rate <strong>of</strong> 86.7%.<br />

The 21-day survival rate in the plg –/– group was 50%. However, when 16 times more bacteria were<br />

injected to the mice, the average survival day for the plg –/– mice was 3 days longer than for the<br />

plg +/+ mice. 24 hours after the induction <strong>of</strong> sepsis, serum IL-6 <strong>and</strong> IL-10 levels <strong>and</strong> the bacterial<br />

counts in all detected organs were significantly higher in plg +/+ mice than in plg –/– mice. In plg +/+<br />

mice, blockade <strong>of</strong> IL-6 by intravenous injection <strong>of</strong> anti-IL-6 antibodies significantly prolonged<br />

the onset <strong>of</strong> mortality <strong>and</strong> improved the survival rate during sepsis. During sepsis, both the<br />

phosphorylated <strong>and</strong> total STAT3 protein levels were dramatically lower in the neutrophils <strong>of</strong><br />

plg –/– mice as compared with plg +/+ mice. These data indicate that plasmin plays different roles<br />

during infection <strong>and</strong> sepsis. In the infectious model plasmin has beneficial effects on host antibacterial<br />

immune responses. However, in the septic model, plasmin seems to excessively promote<br />

the production <strong>of</strong> inflammatory cytokines <strong>and</strong> tissue destruction, cripples neutrophil function, as<br />

well as impairs bacterial killing ability in plg +/+ mice.<br />

48 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i037i<br />

Proteolytic <strong>Activation</strong> <strong>of</strong> the Human Urokinase/Plasmin System<br />

by Staphylococcus aureus<br />

Beaufort N* 1 , Wojciechowski P 1,2 , Sommerh<strong>of</strong>f CP 3 , Schmitt M 1 , Potempa J 2 , Magdolen V 1<br />

1 Department <strong>of</strong> Obstetrics <strong>and</strong> Gynecology, Technical University <strong>of</strong> Munich, Munich, Germany;<br />

2Department <strong>of</strong> Microbiology, Faculty <strong>of</strong> Biochemistry, Biophysics <strong>and</strong> Biotechnology, Jagiellonian<br />

University, Krakow, Pol<strong>and</strong>;<br />

3Department <strong>of</strong> Clinical Chemistry <strong>and</strong> Clinical Biochemistry, Ludwig-Maximilians-University,<br />

Munich, Germany<br />

Presenting author e-mail: nbeaufortgbb@yahoo.fr<br />

The major opportunistic human pathogen Staphylococcus aureus is known to interact with <strong>and</strong><br />

engage the human plasminogen activation system through plasmin(ogen) binding <strong>and</strong> nonproteolytic<br />

activation, a capacity which is thought to participate to bacterial spread <strong>and</strong> invasion.<br />

We here report that staphylococcal reference <strong>and</strong> clinical strains efficiently convert the zymogen<br />

pro-urokinase (pro-uPA) into its active counterpart uPA. Using selective protease inhibitors,<br />

purified bacterial proteases, <strong>and</strong>/or protease-deficient bacterial strains, we establish that this<br />

processing involves the secreted thermolysin-like metalloprotease aureolysin, <strong>and</strong> displays a<br />

2.6 x 10-3 M-1 s-1 catalytic efficiency. Additionally, aureolysin not only targets pro-uPA, but also<br />

cleaves <strong>and</strong> disables the serpins plasminogen activator inhibitor-1 <strong>and</strong> a2-antiplasmin, whereas it<br />

converts plasminogen into angiostatin <strong>and</strong> mini-plasminogen.<br />

Altogether, we propose that, in parallel to the staphylokinase-dependent activation <strong>of</strong><br />

plasminogen, aureolysin contributes significantly to the activation <strong>of</strong> the fibrinolytic system by S.<br />

aureus, <strong>and</strong> thus may promote bacterial spread <strong>and</strong> invasion.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 49


i038i<br />

The Maintenance <strong>of</strong> High Affinity <strong>Plasminogen</strong> Binding by PAM<br />

Variants from Group A Streptococci Is Mediated by Conserved<br />

Arg <strong>and</strong> His Residues in Both the a1 <strong>and</strong> a2 Repeat Domains<br />

Ranson M* 1 , S<strong>and</strong>erson-Smith ML 1 , Walker MJ 1 , Fu Q 2 , Castellino FJ 2 , Prorok M 2<br />

1 School <strong>of</strong> Biological Sciences, University <strong>of</strong> Wollongong, Australia;<br />

2W.M. Keck Center for Transgene Research <strong>and</strong> Department <strong>of</strong> Chemistry <strong>and</strong> Biochemistry, University<br />

<strong>of</strong> Notre Dame, Notre Dame, Indiana, USA<br />

Presenting author e-mail: mranson@uow.edu.au<br />

Subversion <strong>of</strong> the plasminogen activation system is implicated in the virulence <strong>of</strong> Group A<br />

streptococci (GAS). GAS display several receptors for human plasminogen on the cell surface<br />

including the high affinity plasminogen-binding group A streptococcal M protein (PAM),<br />

which mediates an important virulence mechanism for a subset <strong>of</strong> GAS isolates. The major<br />

plasmin(ogen)-binding domain <strong>of</strong> PAM is comprised <strong>of</strong> two characteristic t<strong>and</strong>em repeats<br />

designated a1 <strong>and</strong> a2. We have recently shown that mutation <strong>of</strong> previously identified key<br />

residues Lys98, Arg101, His102 <strong>and</strong> Lys111 within the a1 <strong>and</strong> a2 repeats reduced, but did not<br />

abrogate plasminogen binding by full-length PAM using solid-phase binding assays. Loss <strong>of</strong><br />

plasminogen binding was only observed following simultaneous mutation <strong>of</strong> Arg101, Arg114,<br />

His102 <strong>and</strong> His115 in both the a1 <strong>and</strong> a2 repeats regardless <strong>of</strong> the presence <strong>of</strong> residues Lys98<br />

<strong>and</strong> Lys111. The on- <strong>and</strong> <strong>of</strong>f-rate constants underlying the Kd values obtained from these assays<br />

were further examined using PAM <strong>and</strong> select variants by surface plasmon resonance (SPR). The<br />

relative affinity <strong>of</strong> PAM for streptokinase-resistant murine plasminogen compared to human<br />

plasminogen was also investigated by SPR <strong>and</strong> will be reported. This demonstration <strong>of</strong> a nonlysine-dependent,<br />

high affinity interaction between plasminogen <strong>and</strong> a receptor is unique. We<br />

suggest that the highly conserved Arg <strong>and</strong> His residues from either the a1 or a2 repeats in PAMlike<br />

proteins compensates for variation elsewhere in the binding repeats, or indeed for loss <strong>of</strong> one<br />

<strong>of</strong> the repeats, <strong>and</strong> explains the maintenance <strong>of</strong> high affinity plasminogen binding by naturally<br />

occurring PAM-like variants.<br />

50 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i039i<br />

<strong>Plasminogen</strong> Activator Inhibitor-1 (PAI-1) Is an<br />

Inhibitor <strong>of</strong> Factor VII-Activating Protease in Patients<br />

with Acute Respiratory Distress Syndrome<br />

Wygrecka M* 1 , Morty RE 2 , Markart P 2 , Kanse SM 1 , Andreasen PA 3 , Wind T 3 , Guenther A 2 , Preissner KT 1<br />

1 2 Department <strong>of</strong> Biochemistry, Department <strong>of</strong> Internal Medicine, Faculty <strong>of</strong> Medicine, Justus-Liebig-<br />

University Giessen, Giessen, Germany;<br />

3 Department <strong>of</strong> <strong>Molecular</strong> <strong>and</strong> Structural <strong>Biology</strong>, University <strong>of</strong> Aarhus, Aarhus, Denmark<br />

Presenting author e-mail:malgorzata.wygrecka@innere.med.uni-giessen.de<br />

Factor VII-activating protease (FSAP) is a novel plasma-derived serine protease structurally<br />

homologous to tissue-type (tPA) <strong>and</strong> urokinase-type (uPA) plasminogen activators. We<br />

demonstrate that plasminogen activator inhibitor-1 (PAI-1), the predominant inhibitor <strong>of</strong> tPA <strong>and</strong><br />

uPA in plasma <strong>and</strong> tissues, is an inhibitor <strong>of</strong> FSAP as well. We detected PAI-1-FSAP complexes<br />

in addition to high levels <strong>of</strong> extracellular RNA, an important FSAP c<strong>of</strong>actor, in bronchoalveolar<br />

lavage fluids from patients with acute respiratory distress syndrome (ARDS). Hydrolytic activity<br />

<strong>of</strong> FSAP was inhibited by PAI-1 with a second-order inhibition rate constant (Ka) <strong>of</strong> 3.38 ± 1.12 x<br />

10 5 M-1.s-1. Residue Arg346 was a critical recognition element on PAI-1 for interaction with FSAP.<br />

RNA, but not DNA, fragments (>400 nucleotides in length) dramatically enhanced the reactivity<br />

<strong>of</strong> PAI-1 with FSAP, <strong>and</strong> 4 µg.ml-1 RNA increased the Ka to 1.61 ± 0.94 x 10 6 M-1.s-1. RNA also<br />

stabilized the active conformation <strong>of</strong> PAI-1, increasing the half-life for spontaneous conversion <strong>of</strong><br />

active to latent PAI-1 from 48.4 ± 8 min to 114.6 ± 5 min. In contrast, little effect <strong>of</strong> DNA on PAI-1<br />

stability was apparent. Residues Arg76 <strong>and</strong> Lys80 in PAI-1 were key elements mediating binding<br />

<strong>of</strong> nucleic acids to PAI-1. FSAP-driven inhibition <strong>of</strong> vascular smooth muscle cell proliferation<br />

was antagonized by PAI-1, suggesting functional consequences for the FSAP-PAI-1 interaction.<br />

These data indicate that extracellular RNA <strong>and</strong> PAI-1 can regulate FSAP activity, thereby playing<br />

a potentially important role in hemostasis <strong>and</strong> cell functions under various pathophysiological<br />

conditions such as ARDS.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 51


i040i<br />

Improved Muscle Regeneration in PAI-1-deficient Mice<br />

Is Associated with an Enhanced Inflammatory Response<br />

<strong>and</strong> Reduced Fibrin Deposition after Injury<br />

Ardite E*, Vidal B, Jardí M, González B, Muñoz-Cánoves P<br />

Center for Genomic Regulation (CRG), Program on Differentiation <strong>and</strong> Cancer, Barcelona, Spain<br />

Presenting author e-mail: esther.ardite@crg.es<br />

We have previously shown that uPA activity is required for muscle regeneration in vivo, since<br />

muscle regeneration was impaired in uPA-deficient mice (uPA–/–) after injury, correlating with<br />

persistent accumulation <strong>of</strong> fibrin. In contrast, mice lacking the uPA inhibitor PAI-1 (PAI-1–/–)<br />

showed an improved <strong>and</strong> accelerated muscle regeneration with respect to wild-type (WT) mice<br />

after injury, which correlated with decreased fibrin deposition in PAI-1–/– muscles. These data<br />

suggest that the regulation <strong>of</strong> fibrinolysis by the uPA/PAI-1 balance is important for efficient<br />

muscle regeneration. Moreover, while the recruitment <strong>of</strong> inflammatory cells to regenerating<br />

muscle was reduced in uPA–/– mice, it was increased in PAI-1–/– mice with respect to WT<br />

mice, suggesting that uPA/PAI-1 activities may play a role in the inflammatory response at the<br />

site <strong>of</strong> injury. To test this hypothesis, we analyzed whether transplantation <strong>of</strong> WT bone marrow<br />

would rescue the defective regeneration <strong>of</strong> uPA–/– mice. Transplantation <strong>of</strong> WT bone marrow<br />

into uPA–/– mice restored the inflammatory cell recruitment to damaged uPA–/– muscle to the<br />

levels normally observed in WT mice; more importantly, this normalized inflammatory response<br />

was accompanied by a rescued muscle regeneration in uPA–/– mice. Additional transplantation<br />

experiments <strong>of</strong> WT bone marrow into PAI-1 –/– mice are ongoing. Altogether, these results<br />

indicate that the uPA/PAI-1 balance may be important for efficient inflammation <strong>and</strong> muscle<br />

regeneration, with important implications for human muscular dystrophies.<br />

52 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i041i<br />

Invasion <strong>and</strong> Metastasis <strong>of</strong> Carcinoma Cells Is Prevented by<br />

Urokinase-Derived Antagonists <strong>of</strong> avb5 Integrin <strong>Activation</strong><br />

Franco P 1 , Vocca I 1 , Alfano D 1 , Votta G 1 , Carriero MV 2 , Estrada Y 4 , Netti PA 3 , Ossowski L 4 ,<br />

Stoppelli MP* 1<br />

1 Institute <strong>of</strong> Genetics <strong>and</strong> Biophysics ‘’Adriano Buzzati-Traverso’’, National Research Council, Naples,<br />

Italy;<br />

2 National Cancer Institute <strong>of</strong> Naples, Via M. Semmola, Naples, Italy;<br />

3 Department <strong>of</strong> Materials <strong>and</strong> Production Engineering, University ‘’Federico II’’, Piazzale Tecchio,<br />

Naples, Italy;<br />

4 Department <strong>of</strong> Medicine, Mount Sinai School <strong>of</strong> Medicine, New York, New York, USA<br />

Presenting author e-mail: stoppell@igb.cnr.it<br />

We have previously shown that phosphorylation <strong>of</strong> urokinase (uPA) as well as substitution <strong>of</strong><br />

the critical serine with glutamic acid residues (His-uPA138E/303E) impairs its motogen ability.<br />

Phospho-mimicking uPA strongly inhibited carcinoma cell invasion, as shown by the reduced<br />

metastatic ability <strong>of</strong> HEp3 cell clones stably expressing His-uPA138E/303E in the chick embryo<br />

chorioallantoic membrane. The relationship between this inhibitory effect <strong>and</strong> urokinase receptor<br />

(uPAR) expression was analyzed by using several uPA variants lacking the uPAR binding<br />

domain (residues 9-45 <strong>of</strong> the human sequence) <strong>and</strong> carrying the relevant Ser to Glu substitutions<br />

(dGFa138E/303E) in HEK-293 cells. A strong inhibitory effect <strong>of</strong> cell migration toward different<br />

matrix chemoattractants resulted, in a growth factor domain- <strong>and</strong> uPAR-independent manner.<br />

The inhibitory uPA reduces the speed <strong>of</strong> cell translocation by 5 fold <strong>and</strong> alters the pattern <strong>of</strong> Factin<br />

formation in HEK-293 cells. His-uPA138E/303E <strong>and</strong> dGFa138E/303E specifically bind to<br />

avb5 integrin, as binding is inhibited by specific anti-integrin antibodies <strong>and</strong> by pre-incubation<br />

<strong>of</strong> lig<strong>and</strong> with purified avb5. Finally, cell exposure to picomolar concentrations <strong>of</strong> His-uPA138E/<br />

303E or dGFa138E/303E decreases the affinity <strong>of</strong> avb5 integrin for 125I-vitronectin <strong>and</strong> impairs<br />

cell chemotactic response to vitronectin. Also, the extent <strong>of</strong> talin associated to h5 is markedly<br />

decreased. Although the uPA-derived antagonists do not bind directly to a5b1, a4b1, a1b1,<br />

they prevent cell response to fibronectin, laminin <strong>and</strong> collagen. Similarly, a general inhibition is<br />

observed in the presence <strong>of</strong> the natural antagonist endostatin, which binds to avb5 <strong>and</strong> in the<br />

presence <strong>of</strong> anti- avb5 blocking antibodies. These data, taken together, show that avb5 can convey<br />

a strong signal inhibiting cell migration <strong>and</strong> suggest that this inhibition extends to other integrins.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 53


i042i<br />

Urokinase Receptor/Integrin Interactions in Lung Tumor Development<br />

Tang CH, Hill M, Kim Y, Wei Y, Chapman HA*<br />

Pulmonary <strong>and</strong> Critical Care Division, Department <strong>of</strong> Medicine, Cardiovascular Research Institute,<br />

University <strong>of</strong> California, San Francisco, California, USA<br />

Presenting author e-mail: hal.chapman@ucsf.edu<br />

Upregulation <strong>of</strong> urokinase receptors (uPAR) is common during cancer progression <strong>and</strong> thought to<br />

promote invasion <strong>and</strong> metastasis. Urokinase receptors bind urokinase <strong>and</strong> a set <strong>of</strong> beta1 integrins,<br />

but it remains unclear whether or how uPAR/integrin binding contributes to tumor progression.<br />

Using site-directed mutagenesis, single amino acid mutants <strong>of</strong> the urokinase receptor were<br />

identified that fail to associate with either integrins alpha3beta1 (D262A) or alpha5beta1 (H249A).<br />

A recombinant uPAR bearing both D262A <strong>and</strong> H249A mutations (H/D uPAR) was found to be<br />

expressible, to bind urokinase normally, but to no longer co-precipitate with beta1 integrins. To<br />

study the functional effects <strong>of</strong> these mutations, endogenous uPAR was first stably silenced in<br />

H1299 lung cancer cells <strong>and</strong> then wild type or mutant uPARs expressed. Orthotopic implantation<br />

into nude mice <strong>of</strong> H1299 cells with uPAR knockdown resulted in lower incidence <strong>and</strong> markedly<br />

suppressed lung tumor area (8 mm2 vs 1.5 mm2 , n = 14) compared to parent cells at 32 days,<br />

<strong>and</strong> this was reversed by wild type but not H/D uPAR re-constitution. H1299 cells expressing<br />

wt <strong>and</strong> H/D mutant uPAR were compared for matrix-induced ERK activation, MMP expression,<br />

<strong>and</strong> invasion through matrigel-coated transwells in vitro. Compared with wt uPAR reconstituted<br />

H1299 cells, H/D uPAR cells showed virtually no fibronectin or laminin-induced ERK activation,<br />

MMP upregulation, or matrix-dependent migration. Collectively, these data indicate that uPAR/<br />

integrin complexes are required for matrix-dependent induction <strong>of</strong> MMP through an ERKdependent<br />

pathway in H1299 cells <strong>and</strong> that this signaling pathway appears important to lung<br />

tumor development in vivo.<br />

54 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i043i<br />

uPA- <strong>and</strong> uPAR-Expressing Stromal Cells Accompany<br />

the Transition to Invasive Breast Cancer<br />

Nielsen BS* 1 , Rank F 2 , Illemann M 1 , Lund LR 1 , Danø K 1<br />

1 The Finsen Laboratory, Rigshospitalet, Copenhagen, Denmark;<br />

2 Department <strong>of</strong> Pathology, Rigshospitalet, Copenhagen, Denmark<br />

Presenting author e-mail: schnack@finsenlab.dk<br />

The transition from ductal carcinoma in situ (DCIS) <strong>of</strong> the breast to invasive ductal carcinoma is<br />

facilitated by proteolytic degradation <strong>of</strong> the basement membrane. The transition can be identified<br />

as microinvasive foci in a small proportion <strong>of</strong> DCIS lesions. We have previously reported that<br />

MMP-13 is frequently expressed in such foci. To establish if plasmin-directed proteolysis is<br />

likely to be involved in early invasion we have studied the expression <strong>of</strong> urokinase plasminogen<br />

activator (uPA) <strong>and</strong> its receptor (uPAR) in human DCIS lesions with <strong>and</strong> without microinvasion.<br />

uPA mRNA was detected in periductal stromal cells in all <strong>of</strong> 9 DCIS lesions with microinvasion<br />

<strong>and</strong> in 2 <strong>of</strong> 9 DCIS lesions without microinvasion by in situ hybridization. The uPA mRNA<br />

signal was seen in numerous stromal cells in microinvasive areas together with MMP-13 mRNA<br />

expressing cells. Double immun<strong>of</strong>luorescence analyses, using emission fingerprinting, showed<br />

that the uPA expressing stromal cells included both macrophages <strong>and</strong> my<strong>of</strong>ibroblasts. uPAR was<br />

focally upregulated in periductal stromal cells in all <strong>of</strong> the 9 DCIS lesions with microinvasion<br />

<strong>and</strong> in only 2 <strong>of</strong> the 9 DCIS lesions without microinvasion. uPAR was seen in both macrophages<br />

<strong>and</strong> my<strong>of</strong>ibroblasts in microinvasive areas <strong>and</strong> it was evident that uPA <strong>and</strong> uPAR co-localized<br />

in both fibroblast-like cells <strong>and</strong> macrophage-like cells. The early invasive carcinoma cells were<br />

negative for both uPA <strong>and</strong> uPAR. We conclude that periductal macrophages <strong>and</strong> my<strong>of</strong>ibroblasts<br />

are strongly involved in the initial steps <strong>of</strong> breast cancer invasion by focally upregulating the<br />

expression <strong>of</strong> the plasminogen activation system <strong>and</strong> MMP-13.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 55


i044i<br />

Generation <strong>of</strong> the Malignant Phenotype in HT-1080 Tumor Cells by<br />

PAI-1 Involves Modulation <strong>of</strong> Proteasomal Activity <strong>and</strong> Phosphatases<br />

Mihaly J* 1 , Carroll VA 2 , Breuss JM 1 , Prager GW 1,3 , Binder BR 1<br />

1 Department <strong>of</strong> Vascular <strong>Biology</strong> <strong>and</strong> Thrombosis Research, Medical University Vienna, Austria;<br />

2 Imperial College <strong>of</strong> Medicine, Oxford, United Kingdom;<br />

3 First Clinic for Internal Medicine, Oncology, Medical University Vienna, Austria<br />

Presenting author e-mail: judit.mihaly@meduniwien.ac.at<br />

Elevated levels <strong>of</strong> PAI-1 are <strong>of</strong>ten found in several types <strong>of</strong> cancer <strong>and</strong> increased PAI-1 levels<br />

positively correlate with poor outcome <strong>and</strong> prognosis. We described that prolonged exposure <strong>of</strong><br />

‘’non-malignant’’, low PAI-1 secreting HT-1080 cells to exogenous PAI-1 increased cell adhesion<br />

via redistribution <strong>of</strong> integrins <strong>of</strong> the broad substrate specificity alpha-v heterodimer type to<br />

the cell surface. This PAI-1 induced integrin redistribution was dependent on the presence <strong>of</strong><br />

uPA, uPAR <strong>and</strong> a member <strong>of</strong> the LDLR-family as shown by the use <strong>of</strong> firboblasts derived from<br />

respective gene deficient mice <strong>and</strong> involved sustained activation <strong>of</strong> the ERK/MAPK pathway.<br />

The involvement <strong>of</strong> ERK1 in that process was also demonstrated by overexpression <strong>of</strong> EGFP<br />

tagged ERK1 that mimicked integrin redistribution <strong>and</strong> cell adhesion. Furthermore, exposure<br />

to exogenous PAI-1 also influences proteasomal activity: at early timepoints (2 hours) a PAI-1/<br />

uPA/LDLR dependent increase <strong>of</strong> proteasomal activity was found, whereas after 18 hours the<br />

proteasomal activity in PAI-1 stimulated cells dropped below levels measured in unstimulated<br />

control. The activity <strong>of</strong> ERK1/2 is tightly controlled by dual specificity phosphatases, whereby<br />

one <strong>of</strong> the major players is the MKP1/DUSP1 <strong>and</strong> exposure <strong>of</strong> cells to exogenous PAI-1 also<br />

modulated DUSP-1 on the mRNA <strong>and</strong> protein levels, while DUSP4 remained unaffected. Active<br />

ERK1/2 controls proteasomal degradation <strong>of</strong> its regulator MKP1/DUSP1 by phosphorylation.<br />

From these data we propose a model in which PAI-1 triggered sustained ERK activation is<br />

maintained by a fine tuned regulatory mechanism involving modulation <strong>of</strong> the proteasomal<br />

activity providing the optimal kinase-phosphatase balance required for survival <strong>of</strong> detached/<br />

metastasizing <strong>and</strong> proliferation <strong>of</strong> the lodged cells.<br />

56 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i045i<br />

PEGylated DX-1000: Pharmacokinetics: Anti-Tumor <strong>and</strong><br />

Anti-Metastatic Effects <strong>of</strong> a Specific Plasmin Inhibitor<br />

Devy L* 1 , Rabbani SA 2 , Stochl M 3 , Ruskowski M 4 , Mackie I 5 , Naa L 1 , Toews M 3 , van Gool R 1 , Chen J 3 ,<br />

Ley A 3 , Ladner RC 3 , Dransfield DT 3 , Henderikx P 1<br />

1 Dyax s.a. Building 22, Sart-Tilman, Liege, Belgium;<br />

2 Department <strong>of</strong> Medicine <strong>and</strong> Oncology, McGill University Health Centre, Montreal, QC, Canada;<br />

3 Dyax Corporation, Cambridge, Massachusetts, USA;<br />

4Department <strong>of</strong> Radiology, Division <strong>of</strong> Nuclear Medicine, University <strong>of</strong> Massachusetts Medical School,<br />

Worcester, Massachusetts, USA;<br />

5 Haematology Department, University College London, London, United Kingdom<br />

Presenting author e-mail: ldevy@dyax.com<br />

Using phage display, we identified a TFPI-derived Kunitz domain protein which is a specific high<br />

affinity inhibitor <strong>of</strong> plasmin (Ki=99±15pM) referred to as DX-1000. DX-1000 blocked plasminmediated<br />

proMMP-9 activation, invasiveness (44-66%) <strong>of</strong> uPA-expressing HT-1080 cells <strong>and</strong><br />

tube formation <strong>of</strong> HUVECs (IC 50 =1.4±0.3nM) <strong>and</strong> mouse endothelial cells (IC 50 =16.6± 0.1nM).<br />

DX-1000 did not significantly affect haemostasis <strong>and</strong> coagulation in vitro. However, due to its<br />

low molecular weight (~7 kDa), the protein exhibited a rapid plasma clearance rate in vivo (b<br />

phase half-life=27 minutes in mice <strong>and</strong> 1 hour in rabbits). After site-specific PEGylation, DX-1000<br />

retained its activity <strong>and</strong> exhibited a decreased plasma clearance (b phase half-life=13 hours in<br />

mice <strong>and</strong> 59 hours in rabbits). 4PEG-DX-1000 was effective in vitro, <strong>and</strong> inhibited growth <strong>of</strong> MDA-<br />

MB231 tumor cells in nude mice (45% TGI). 4PEG-DX-1000 treatment caused a marked inhibition<br />

<strong>of</strong> angiogenesis, a significant reduction <strong>of</strong> uPA expression, <strong>and</strong> inhibition <strong>of</strong> tumor proliferation<br />

<strong>and</strong> MAPK phosphorylation. 4PEG-DX-1000 treatment significantly reduced the number <strong>of</strong><br />

metastatic foci in the lung (46%) <strong>and</strong> the liver (57%) <strong>and</strong> decreased the level <strong>of</strong> active FAK in the<br />

primary tumors. Together, the results from these studies provide compelling evidence for the role<br />

<strong>of</strong> plasmin inhibitors as therapeutic agents for blocking breast cancer growth <strong>and</strong> metastasis.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 57


i046i<br />

Cytotoxic Potential <strong>of</strong> a Novel uPA-activity Dependent<br />

<strong>and</strong> EGF Receptor-targeting Pro-drug<br />

Rønø B* 1, 5 , Bae Kim G 2 , Liu S 3 , Kristjansen PEG 4 , Neville DM 2 , Leppla SH 3 , Bugge TH 5 , Rømer J 1<br />

1 Finsen Laboratory, Rigshospitalet, Copenhagen, Copenhagen, Denmark;<br />

2 National Institutes <strong>of</strong> Mental Health, NIH, Bethesda, Maryl<strong>and</strong>, USA;<br />

3 National Institutes <strong>of</strong> Allergy <strong>and</strong> Infectious Diseases, NIH, Bethesda, Maryl<strong>and</strong>, USA;<br />

4 Institute <strong>of</strong> <strong>Molecular</strong> Pathology, University <strong>of</strong> Copenhagen, Copenhagen, Denmark;<br />

5 National Institutes <strong>of</strong> Dental <strong>and</strong> Crani<strong>of</strong>acial Research, NIH, Bethesda, Maryl<strong>and</strong>, USA<br />

Presenting author e-mail: brono@finsenlab.dk<br />

To exploit this preferential expression <strong>of</strong> uPA <strong>and</strong> uPAR in cancer we have constructed a novel<br />

pro-drug whose activation is dependent on proteolytic processing by uPA. To further improve<br />

the cancer specificity, the pro-drug targets epidermal growth factor receptor (EGFR) expressing<br />

cells solely. The pro-drug, termed DT-U2-TGFa, was generated by replacing the native proteolytic<br />

activation sequence in the diphtheria toxin with an amino acid sequence recognised by uPA <strong>and</strong><br />

in addition by exchanging the diphtheria receptor-binding domain with transforming growth<br />

factor alpha (TGFa).<br />

We have examined the activation, specificity, <strong>and</strong> cytotoxic potential <strong>of</strong> the pro-drug. By<br />

incubating the pro-drug with recombinant uPA DTU2TGFa was cleavaged by uPA in a<br />

time dependent manner. In a cell culture based assay, the uPA/uPAR <strong>and</strong> EGFR expressing<br />

human head <strong>and</strong> neck cancer cell line, HN6, bound, activated, <strong>and</strong> internalised DT-U2-TGFa.<br />

Furthermore, a colorimetric cytotoxicity assay showed that the cells were killed by DT-U2-TGFa<br />

in a dose dependent manner. Incubation with different uPA inhibitors prior to treatment with<br />

DT-U2-TGFa efficiently reduced activation <strong>of</strong> the pro-drug, substantiating the requirement for<br />

active uPA. EGFR transduced murine fibroblasts, NR6W, were sensitive to DT-U2-TGFa, while<br />

the parental EGFR negative NR6 cells were not affected, indicating that the cytotoxicity <strong>of</strong> DTU2-<br />

TGFa is restricted to EGFR expressing cells. In support <strong>of</strong> these data HN6 cells pre-incubated with<br />

TGFa were completely rescued from DT-U2-TGFa induced cell death.<br />

In conclusion, we have constructed, produced, <strong>and</strong> tested a novel pro-drug that targets uPA/<br />

uPAR <strong>and</strong> EGFR expressing cells. Our data provides evidence <strong>of</strong> a highly specific pro-drug with<br />

promising cytotoxic efficacy in cancer cells.<br />

58 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i047i<br />

Inhibition <strong>of</strong> Mouse uPA Activity by Mouse<br />

Monoclonal Antibodies in vitro <strong>and</strong> in vivo<br />

Lund IK*, Jögi A, Behrendt N, Ploug M, Gårdsvoll H, Lund LR, Rømer J, Høyer-Hansen G<br />

Finsenlaboratoriet, Copenhagen BioCenter, Copenhagen, Denmark<br />

Presenting author e-mail: ikl@finsenlab.dk<br />

The potential <strong>of</strong> monoclonal antibodies (mAbs) as anti-cancer therapeutics has lately been<br />

demonstrated for a number <strong>of</strong> antigens. To enable in vivo therapy experiments using mAbs in<br />

murine cancer models, we immunized mice deficient in murine urokinase plasminogen activator<br />

(muPA) with recombinant muPA, thus developing murine mAbs directed against muPA.<br />

We have selected 5 mAbs (mU1-mU5) reacting with muPA in ELISA, Surface Plasmon Resonance<br />

(SPR) analysis, <strong>and</strong> Western blotting. Analysing the epitope location on muPA using the<br />

recombinant amino-terminal fragment <strong>of</strong> muPA (mATF) <strong>and</strong> the B-chain revealed that only<br />

mU1 recognized an epitope exclusively located in the B-chain, encompassing the catalytic site <strong>of</strong><br />

muPA. SPR analyses demonstrated that mU2 was unable to bind receptor (muPAR)-bound muPA<br />

<strong>and</strong> prevented binding <strong>of</strong> muPAR to muPA. In vitro cell binding experiments using 125I-mATF<br />

illustrated an efficient mU2-induced interference <strong>of</strong> muPA-muPAR interaction on the cell surface.<br />

Using an enzyme kinetic assay measuring muPA-dependent plasminogen activation, four mAbs<br />

were found to inhibit muPA catalytically activity to various extend, with mU1 having the most<br />

pronounced effect. Application <strong>of</strong> mU1 or mU2 in an anthrax toxin assay dependent on muPA<br />

activity yielded a very high cell rescue.<br />

Importantly, treatment <strong>of</strong> tissue-type plasminogen (tPA) deficient mice with mU1 resulted in<br />

significantly delayed wound healing mimicking the phenotype observed in uPA;tPA doubledeficient<br />

mice. Moreover, data concerning the in vivo effect <strong>of</strong> mU1 on liver fibrin deposits <strong>of</strong> tPA<br />

deficient mice as compared to uPA;tPA double-deficient mice will be provided.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 59


i048i<br />

Crystal Structure <strong>of</strong> Human Urokinase Complexed<br />

with a Cyclic Peptidyl Inhibitor, uPAin-1<br />

Zhao G 1 , Yuan C 1 , Bian C 1 , Wind T 2 , Andreasen PA 2 , Huang M* 1<br />

1State Key Laboratory <strong>of</strong> Structural Chemistry, Fujian Institute <strong>of</strong> Research on the Structure <strong>of</strong> Matter,<br />

Chinese Academy <strong>of</strong> Sciences, Fuzhou, Fujian, China;<br />

2 Department <strong>of</strong> <strong>Molecular</strong> <strong>Biology</strong>, University <strong>of</strong> Aarhus, Aarhus, Denmark<br />

Presenting author e-mail: mhuang@fjirsm.ac.cn<br />

Urokinase-type plasminogen activator (uPA) plays a crucial role in the regulation <strong>of</strong> tumor<br />

cell adhesion <strong>and</strong> migration. The inhibition <strong>of</strong> uPA activity is a promising mechanism for anticancer<br />

therapy. A cyclic peptidyl inhibitor, upain-1, CSWRGLENHRMC, was identified recently<br />

as a competitive <strong>and</strong> highly specific uPA inhibitor (Hansen et al. J Biol Chem 280, 38424-37). We<br />

determined the crystal structure <strong>of</strong> uPA in complex with upain-1 at 2.15 Å. The structure reveals<br />

that the cyclic peptide adopts a rigid conformation stabilized by two tight beta turns formed by<br />

Leu6-His9 <strong>and</strong> His9-Cys12 segments, respectively. The Glu7 residue <strong>of</strong> upain-1 forms hydrogen<br />

bonds with the main chain nitrogen atoms <strong>of</strong> residues 4, 5, <strong>and</strong> 6 <strong>of</strong> upain-1, <strong>and</strong> is also critical<br />

for maintaining the active conformation <strong>of</strong> upain-1. The Arg4 <strong>of</strong> upain-1 is inserted into the uPA’s<br />

specific S1 pocket. The Ser2 residue <strong>of</strong> upain-1 locates close to the S1b pocket <strong>of</strong> uPA. The Gly5<br />

<strong>and</strong> Glu7 residues <strong>of</strong> upain-1 occupy the S2 pocket <strong>and</strong> the oxyanion hole <strong>of</strong> uPA, respectively.<br />

Furthermore, the Asn8 residue <strong>of</strong> upain-1 binds to the 37-loop <strong>and</strong> 60-loop <strong>of</strong> uPA <strong>and</strong> renders<br />

the specificity <strong>of</strong> upain-1 for uPA. The steric hindrance <strong>of</strong> the side chain <strong>of</strong> Glu7 <strong>and</strong> the indole<br />

ring <strong>of</strong> Trp3 residue <strong>of</strong> upain-1 prevents the carboxyl group <strong>of</strong> Arg4 residue at the cleavage site to<br />

be accessible by the catalytic Ser195 residue <strong>of</strong> uPA, thus making upain-1 behave as an inhibitor<br />

rather than a substrate. Based on this structure, we propose a new pharmacophore for the design<br />

<strong>of</strong> highly specific uPA inhibitors.<br />

60 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i049i<br />

Discovery <strong>of</strong> a Novel Zymogen Targeting Inhibitor <strong>of</strong><br />

Urokinase-type <strong>Plasminogen</strong> Activator: Evidence for<br />

Structural Flexibility <strong>of</strong> the Protease Domain<br />

Blouse GE 1 , Bøtkjær KA 1 , Deryugina EI 3 , Kjelgaard S 1 , Byszuk O 1 , Mortensen KK 1 , Quigley JP 3 ,<br />

Andreasen PA* 1<br />

1 2 Laboratory <strong>of</strong> <strong>Cellular</strong> Protein Science, BioDesign Laboratory, Department <strong>of</strong> <strong>Molecular</strong> <strong>Biology</strong>,<br />

University <strong>of</strong> Aarhus, Aarhus, Denmark;<br />

3 Department <strong>of</strong> Cell <strong>Biology</strong>, Scripps Research Institute, La Jolla, California, USA<br />

Presenting author e-mail: pa@mb.au.dk<br />

Considerable interest is emerging for pharmacologic interference with proteolytic enzyme targets.<br />

One such validated target is urokinase-type plasminogen activator (uPA), which has received<br />

much attention due to a suspected causal role in the clinical progression <strong>of</strong> cancer <strong>and</strong> other<br />

processes <strong>of</strong> pathological tissue remodeling. Nonetheless, development <strong>of</strong> specific small molecule<br />

inhibitors has proven a daunting task. In nature, a key mechanism <strong>of</strong> protease regulation is the<br />

point <strong>of</strong> zymogen activation. Thus far, controlling protease activity by targeting the zymogen<br />

activation step is an underexploited strategy. We have now designed a specific monoclonal<br />

antibody inhibitor (Mab-112) with sub-nanomolar affinity to pro-uPA. A detailed mechanistic<br />

evaluation with several biophysical methods elucidated a novel multifunctional mechanism<br />

whereby Mab-112 retards the protelytic conversion <strong>of</strong> single-chain pro-uPA into the two-chain<br />

form <strong>and</strong> subsequently averts the conformational transition to a mature protease by sequestering<br />

the two-chain form in a zymogen-like state. Furthermore, Mab-112 is a non-competitive inhibitor<br />

<strong>of</strong> two-chain uPA, stabilising the protease in a non-catalytic conformation. Functional studies<br />

employing high intravasating (HT-hi/diss) <strong>and</strong> low intravasating (HT-lo/diss) variants <strong>of</strong> the<br />

human HT-1080 cell line demonstrate that Mab-112 is an effective inhibitor <strong>of</strong> intravasation in the<br />

chorioallantoic membrane assay. These findings show the utility <strong>of</strong> pharmacological interference<br />

<strong>of</strong> zymogen activation as a plausible <strong>and</strong> robust means to regulate uPA activity <strong>and</strong> the<br />

downstream effects <strong>of</strong> plasminogen activation. Furthermore, a strategy that targets regions related<br />

to pro-enzyme activation likely provide a unique inhibitor-protease interaction surface <strong>and</strong> is thus<br />

expected to enhance the chances <strong>of</strong> engineering high inhibitor specificity.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 61


i050i<br />

A Novel Type <strong>of</strong> Agent Blocking the Association <strong>of</strong> uPA<br />

to its Receptor uPAR: uPA-binding Aptamers<br />

Dupont DM* 1,2 , Madsen JB 1 , Kjems J 1,2 , Andreasen PA 1,2<br />

1 Department <strong>of</strong> <strong>Molecular</strong> <strong>Biology</strong>, 2 iNANO Centre, University <strong>of</strong> Aarhus, Aarhus, Denmark<br />

Presenting author e-mail: dmd@mb.au.dk<br />

Systematic evolution <strong>of</strong> lig<strong>and</strong>s by exponential enrichment (SELEX) is a relatively new approach<br />

for generating agents <strong>of</strong> potential therapeutic <strong>and</strong> diagnostic interest. The technique combines<br />

the ability <strong>of</strong> RNA or DNA oligonucleotides to fold into a variety <strong>of</strong> three-dimensional structures,<br />

with the possibility <strong>of</strong> selecting, from very large pools <strong>of</strong> r<strong>and</strong>om sequences (~10 15 ), the ones<br />

capable <strong>of</strong> binding to a target <strong>of</strong> interest. We have generated a library <strong>of</strong> serum-stable 2´-fluoropyrimidine<br />

modified RNA oligonucletides <strong>and</strong> used it in a SELEX experiment to select sequences,<br />

or so-called aptamers, binding to human urokinase-type plasminogen activator (uPA). As<br />

analysed by surface plasmon resonance (SPR), the aptamers bind to the amino terminal fragment<br />

<strong>of</strong> human, but not murine, uPA, with K D -values in the low nanomolar range. SPR analyses <strong>and</strong><br />

cell binding assays have shown that the aptamers block the association <strong>of</strong> uPA with its receptor<br />

uPAR. In consistency with their binding area in the amino terminal fragment, they do not inhibit<br />

the uPA proteolytic activity directly, but inhibit uPAR-dependent plasminogen activation on cell<br />

surfaces. By RNA sequence alignments <strong>and</strong> computerised secondary structure predictions, we<br />

were able to identify the uPA binding regions <strong>of</strong> the aptamers <strong>and</strong> delete regions unnecessary<br />

for uPA binding. uPA-binding aptamers represent a promising principle for interfering with the<br />

pathophysiological functions <strong>of</strong> the plasminogen activation system. Targeting the receptor-lig<strong>and</strong><br />

interaction through uPA may be advantageous as compared to uPAR-targeting agents which may<br />

not be antagonists <strong>of</strong> all functions <strong>of</strong> uPAR. Aptamers are also potential tools for analytical <strong>and</strong><br />

imaging purposes.<br />

62 X I t h I n t e r n a t i o n a l W o r k s h o p o n


Abstracts for Poster Presentations<br />

i051i<br />

Identification <strong>and</strong> Analysis <strong>of</strong> the Vn Binding Site in Mouse uPAR<br />

Pirazzoli V* 1 , Andolfo AP 1 , Madsen CD 1 , Sidenius N 1,2<br />

1 IFOM Fondation, FIRC Institute <strong>of</strong> <strong>Molecular</strong> Oncology, Milan, Italy;<br />

2 <strong>Molecular</strong> Genetics Unit, DIBIT, Università Vita-Salute San Raffaele, Milan, Italy<br />

Presenting author e-mail: valentina.pirazzoli@ifom-ieo-campus.it<br />

A complete functional alanine scan <strong>of</strong> human uPAR conducted recently in our laboratory<br />

identified the interaction with Vn as being the only essential direct uPAR-interaction required for<br />

the receptor to induce changes in cell adhesion, morphology, migration <strong>and</strong> signaling. To address<br />

the importance <strong>of</strong> the Vn-interaction in vivo it is our goal to generate a knock-in mouse where the<br />

uPAR gene has been modified to encode a receptor deficient in Vn-binding but otherwise normal<br />

(i.e. with normal uPA-binding). With such a goal in mind, we now report the analysis <strong>of</strong> the<br />

Vn-binding epitope in mouse uPAR. When over-expressed in CHO cells, wild-type mouse uPAR<br />

(muPAR) was found to induce changes in cell morphology <strong>and</strong> colony formation very similar<br />

to that <strong>of</strong> the human receptor. These changes include the formation <strong>of</strong> actin-rich lamellipodia,<br />

loss <strong>of</strong> stress fibers, reduced cell-cell contact as well as a complete failure to form colonies when<br />

seeded a low density. We had previously identified 5 residues in human uPAR which appear to<br />

engage Vn directly <strong>and</strong> which when changed into alanine disrupt the Vn-interaction completely<br />

or partially. To establish if the corresponding, fully conserved, residues in muPAR (W32, R58, I63,<br />

R92 <strong>and</strong> Y93) are also responsible for Vn-binding we mutated these into alanine <strong>and</strong> tested their<br />

ability to induce RGD-independent Vn-adhesion as well as changes in cell morphology. The result<br />

<strong>of</strong> this analysis documented that a substitution <strong>of</strong> any <strong>of</strong> these residues completely impaired<br />

muPAR binding to Vn. Additional experiments using recombinant soluble muPAR confirmed the<br />

impaired Vn-binding <strong>and</strong> a normal uPA-binding.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 63


i052i<br />

Novel uPAR Binding Site in Vitronectin<br />

Andolfo A* 1 <strong>and</strong> Sidenius N 1,2<br />

1 FIRC Institute <strong>of</strong> <strong>Molecular</strong> Oncology Foundation, Milan, Italy;<br />

2 Dibit, Fondazione Vita-salute San Raffaele, Milan, Italy<br />

Presenting author e-mail: annapaola.<strong>and</strong>olfo@ifom-ieo-campus.it<br />

The ability <strong>of</strong> the serum protein vitronectin (Vn) to modulate cell adhesion requires the<br />

aminoterminal somatomedin B (SMB) domain (amino acids 1-44) which binds both plasminogen<br />

activator inhibitor-1 (PAI-1) <strong>and</strong> urokinase-type plasminogen activator receptor (uPAR) as well<br />

as the flanking ArgGlyAsp sequence (RGD, amino acids 45-47) responsible for the interaction<br />

with Vn-receptors <strong>of</strong> the integrin family. Here we report our data on uPAR-Vn interaction site<br />

mapping by a complete alanine scan <strong>of</strong> the SMB domain. We expressed <strong>and</strong> purified the mutants<br />

from mammalian cell cultures <strong>and</strong> assayed their biochemical <strong>and</strong> biological activity. In contrast<br />

to previously published data using SMB domains expressed in E. coli, we find that single alanine<br />

substitution only has marginal effects on PAI-1 binding (< 2-fold reduction in affinity) <strong>and</strong> no<br />

detectable effects on integrin mediated RGD-dependent cell adhesion. We confirm the previously<br />

published data on uPAR binding site in Vn, that is overlapping to the known PAI-1 binding site<br />

(F13, D22, L24, Y27, Y28). Most importantly, we find additional residues in the SMB domain that<br />

are exclusively required for uPAR binding (S4, G7, R8), thus defining a composite site with two<br />

distinct epitopes located on opposite sides <strong>of</strong> the SMB-structure.<br />

In conclusion, the presence <strong>of</strong> two separate uPAR binding sites in the SMB, which are both critical<br />

for binding, raises important questions towards how <strong>and</strong> if a single molecule <strong>of</strong> uPAR can engage<br />

these two sites contemporarily.<br />

64 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i053i<br />

Dissecting Serpin-protease Reaction Pathways<br />

by the Use <strong>of</strong> Monoclonal Antibodies<br />

Bødker JS*, Blouse GE, Dupont DM, Andreasen PA<br />

Department <strong>of</strong> <strong>Molecular</strong> <strong>Biology</strong>, University <strong>of</strong> Aarhus, Aarhus, Denmark<br />

Presenting author e-mail: jsb@mb.au.dk<br />

The formation <strong>of</strong> a stable PAI-1-protease complex is believed to progress through several<br />

separate steps. The steps include exosite interactions between the protease 37-loop <strong>and</strong> the serpin<br />

reactive centre loop (RCL); formation <strong>of</strong> a reversible Michaelis complex; formation <strong>of</strong> a stable<br />

acyl-enzyme intermediate; release <strong>of</strong> exosite interactions; insertion <strong>of</strong> the N-terminal side <strong>of</strong> the<br />

RCL as b-str<strong>and</strong> 4A which translocates the protease to the opposite pole <strong>of</strong> the serpin. We have<br />

now obtained further information about the PAI-1-protease mechanism by several approaches<br />

including the use <strong>of</strong> monoclonal anti-PAI-1 antibodies having epitopes near the path believed<br />

to be followed by the protease during translocation, stopped-flow analyses with fluorescently<br />

labeled PAI-1, BIACORE, <strong>and</strong> site-directed mutagenesis. Mab-5, which has an epitope near the<br />

region where the RCL first inserts into b-sheet A, caused a slight reduction in the rate <strong>of</strong> Michaelis<br />

complex formation, but promoted an 100-fold reduction in the rate <strong>of</strong> RCL insertion, an effect<br />

reduced by mutations known to affect the release from the protease-serpin exosite interactions.<br />

Another antibody, Mab-2, with an epitope in a-helix F, caused a 5–10-fold reduction in the rate <strong>of</strong><br />

loop insertion, induction <strong>of</strong> PAI-1 substrate behaviour, <strong>and</strong> accumulation <strong>of</strong> a rapidly dissociating<br />

complex on BIACORE, presumably an acyl-enzyme complex with partial loop insertion <strong>and</strong><br />

lacking exosite interactions. These results lend further support to the existence <strong>of</strong> two forms for<br />

acyl-enzyme intermediates in the serpin-protease reaction mechanism, one with intact exosite<br />

interactions <strong>and</strong> one in which release <strong>of</strong> exosite interactions is coupled to the initial phase <strong>of</strong> RCL<br />

insertion.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 65


i054i<br />

PAI-1-vitronectin Interactions Involve an Extended Binding<br />

Surface <strong>and</strong> Mutual Conformational Rearrangements<br />

Blouse GE 1 , Peterson CB 2 , Dupont DM 1 , Ploug M 3 , Gårdsvoll H 3 , Schar CR 2 , Perron MJ 4 , Minor KH 2 ,<br />

Shore JD 4 , Andreasen PA* 1<br />

1 Department <strong>of</strong> <strong>Molecular</strong> <strong>Biology</strong>, University <strong>of</strong> Aarhus, Aarhus, Denmark;<br />

2Department <strong>of</strong> Biochemistry, <strong>Cellular</strong> <strong>and</strong> <strong>Molecular</strong> <strong>Biology</strong>, University <strong>of</strong> Tennessee, Knoxville,<br />

Tennessee, USA;<br />

3 Finsen Laboratory, Rigshospitalet, Copenhagen, Denmark;<br />

4 Department <strong>of</strong> Pathology, Henry Ford Health System, Detroit, Michigan, USA<br />

Presenting author e-mail: pa@mb.au.dk<br />

It has been proposed that binding <strong>of</strong> PAI-1 to native vitronectin induces a reorganisation <strong>of</strong><br />

vitronectin from a closed to an open conformation allowing assembly <strong>of</strong> vitronectin oligomers<br />

with changed cell <strong>and</strong> matrix binding functions. In order to elucidate early molecular events<br />

during interactions between PAI-1 <strong>and</strong> vitronectin, we have applied a robust strategy that<br />

combines protein engineering, fluorescence spectroscopy <strong>and</strong> rapid reaction kinetics. Fluorescence<br />

stopped-flow experiments indicated a fast, concentration dependent, biphasic binding <strong>of</strong> PAI-<br />

1 to native monomeric vitronectin, but a monophasic association with vitronectins N-terminal<br />

somatomedin B domain (SMB), suggesting that multiple phases <strong>of</strong> the binding interaction<br />

occur only when full-length vitronectin is present. Nonetheless, in all cases, the initial fast<br />

interaction was followed by slower fluorescence changes, which we, by the use <strong>of</strong> an engineered,<br />

fluorescently silent PAI-1 with non-natural amino acids, demonstrated to be caused by reciprocal<br />

structural changes within the PAI-1 structure as well as in native vitronectin. Furthermore,<br />

measuring the effect <strong>of</strong> vitronectin on the rate <strong>of</strong> insertion <strong>of</strong> the reactive centre loop into b-sheet A<br />

<strong>of</strong> PAI-1 during reaction with target proteases by the use <strong>of</strong> fluorescently labelled PAI-1 variants,<br />

we observed that both full-length vitronectin <strong>and</strong> SMB had protease-specific effects on the rate<br />

<strong>of</strong> loop insertion, but that the two had clearly different effects. We interpret these results as<br />

support for a model <strong>of</strong> PAI-1-vitronectin binding in which there is an extended interaction surface<br />

implicating parts <strong>of</strong> vitronectin other than the somatomedin B domain.<br />

66 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i055i<br />

Intact (non-cleavable) Cell-surface u-PAR Accelerates<br />

Clearance <strong>of</strong> tcu-PA:PAI-1:u-PAR Complexes <strong>and</strong> Subsequent<br />

Re-surfacing <strong>of</strong> Intact <strong>and</strong> Functional u-PAR<br />

Nieves-Li EC* <strong>and</strong> Manch<strong>and</strong>a N<br />

Department <strong>of</strong> Biochemistry <strong>and</strong> Internal Medicine, University <strong>of</strong> Illinois at Urbana-Champaign,<br />

Urbana, Illinois, USA<br />

Presenting author e-mail: enieves@uiuc.edu<br />

Urokinase plasminogen activator receptor (u-PAR) has been well-defined as a promoter <strong>of</strong><br />

plasminogen activation <strong>and</strong> modulates several cellular processes such as adhesion, proliferation,<br />

<strong>and</strong> migration. U-PAR has been shown to be susceptible to proteolytic cleavage by several<br />

proteases, one <strong>of</strong> them being its lig<strong>and</strong> two-chain u-PA (tcu-PA). This dissociates D1 from<br />

the rest <strong>of</strong> the anchored receptor <strong>and</strong> leads to loss <strong>of</strong> lig<strong>and</strong> binding. However, exposure <strong>of</strong><br />

the chemotactic epitope after D1 removal gives u-PAR a new role in signaling <strong>and</strong> migration.<br />

We found that cell-surface plasminogen activation is impaired upon D1 removal. We have<br />

investigated the role <strong>of</strong> receptor cleavage in the regulation <strong>of</strong> cell-surface plaminogen activation.<br />

We engineered <strong>and</strong> expressed wt u-PAR as well as tcu-PA cleavage resistant u-PAR (tcr-uPAR)<br />

in HEK293 cells. Using three different approaches to follow u-PAR internalization <strong>and</strong> recycling,<br />

i.e. biotin labeled receptor, biotin-labeled tcu-PA-PAI-1 complexes, <strong>and</strong> assays for plasminogen<br />

activation, we found that intact receptor better promotes the internalization <strong>of</strong> cell-surface u-<br />

PAR:u-PA-PAI-1 complexes <strong>and</strong> promotes subsequent re-surfacing <strong>of</strong> unoccupied <strong>and</strong> functional<br />

receptor. We hypothesize that intact cell-surface u-PAR may accelerate removal <strong>of</strong> inhibited<br />

tcu-PA from the cell-surface environment <strong>and</strong> that its re-expression may promote cell-surface<br />

plasminogen activation.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 67


i056i<br />

The Central b-sheet <strong>of</strong> PAI-1 Demonstrates<br />

Two Dynamically Distinct Regions<br />

Li S* 1 , Lawrence DA 2 , Schwartz BS 1,3<br />

1 Department <strong>of</strong> Biochemistry, University <strong>of</strong> Illinois at Urbana-Chanpaign, Urbana, Illinois, USA;<br />

2 Department <strong>of</strong> Internal Medicine, University <strong>of</strong> Michigan Medical School, Ann Arbor, Michigan, USA;<br />

3 Department <strong>of</strong> Medicine, University <strong>of</strong> Illinois at Urbana-Champaign, Urbana, Illinois, USA<br />

Presenting author e-mail: SLI3@uiuc.edu<br />

PAI-1 is an inhibitory serpin that can spontaneously transit with a half-life <strong>of</strong> 1-2 hours at 37°C<br />

to a latent form whereby the RCL is intact but fully inserted into b-sheet A. The labile nature <strong>of</strong><br />

native PAI-1 is thought to be due to the ability <strong>of</strong> its central sheet to exist in equilibrium between<br />

open <strong>and</strong> closed conformations. The stable variant, 14-1B, contains 4 amino acid substitutions that<br />

convey a 70-fold enhancement in the inhibitory half-life compared to wild-type PAI-1 (wtPAI-<br />

1). X-ray crystal structures <strong>of</strong> latent PAI-1 <strong>and</strong> active 14-1B suggest that some <strong>of</strong> the mutations<br />

in the stable variant keep b-sheet A closed, preventing incorporation <strong>of</strong> the RCL. We tested this<br />

hypothesis using RCL-mimicking peptides, which have been shown to insert into b-sheet A<br />

as s4A. We found that 14-1B is less susceptible than wtPAI-1 to inactivation by such peptides.<br />

The PAI-1 c<strong>of</strong>actor, vitronectin (Vn), did not affect the binding <strong>of</strong> a pentamer, Ac-TVASS, to<br />

14-1B, yet enhanced the binding <strong>of</strong> an octamer, Ac-TVASSSTA. The binding <strong>of</strong> Vn to wtPAI-1<br />

modestly delays the transition to latency. Yet we found that that Vn did not protect wtPAI-1<br />

from inactivation by the peptides, rather slightly enhancing such inactivation. These results are<br />

inconsistent with the currently held model on how active PAI-1 is stabilized by its physiologic<br />

c<strong>of</strong>actor. We propose that sheet A breathes in a pigeon-toed fashion in which the region<br />

underneath a-helix F remains closed <strong>and</strong> the region near the RCL is dynamic.<br />

68 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i057i<br />

Regulation <strong>of</strong> Cancer-Cell Plasmin Generation by<br />

Annexin A2-S100A10 Heterotetramer (AIIt)<br />

Waisman DM*<br />

Department <strong>of</strong> Biochemistry <strong>and</strong> <strong>Molecular</strong> <strong>Biology</strong>, Dalhousie University, Halifax, Nova Scotia,<br />

Canada<br />

Presenting author e-mail: david.waisman@dal.ca<br />

Many findings over several decades strongly suggest an important role for the plasminogen<br />

activation system in cancer cell invasion <strong>and</strong> metastasis. We have investigated the role <strong>of</strong> the<br />

S100A10, annexin A2 <strong>and</strong> the heterotetrameric complex formed by these subunits (AIIt) in<br />

plasminogen regulation. Surface plasmon resonance studies established that S100A10 bound<br />

directly to tPA, plasminogen <strong>and</strong> plasmin whereas the annexin A2 bound only plasmin. AIIt <strong>and</strong><br />

S100A10, but not annexin A2, dramatically stimulated the plasminogen activator-dependent<br />

conversion <strong>of</strong> plasminogen to plasmin in vitro. Immun<strong>of</strong>luorescence microscopy demonstrated<br />

the colocalization <strong>of</strong> S100A10 with the uPA/uPAR complex. We have also utilized both antisense<br />

<strong>and</strong> siRNA methodologies to knockdown S100A10. Loss <strong>of</strong> extracellular S100A10 resulted in a<br />

70-90% loss in cellular plasmin generation. Under these conditions the levels <strong>of</strong> annexin A2 were<br />

unchanged. The S100A10 knockdown cells were also less invasive <strong>and</strong> showed a dramatic loss in<br />

tumor formation <strong>and</strong> metastatic potential. We also observed that the addition <strong>of</strong> plasminogen to<br />

cancer cells resulted in the oxidation <strong>of</strong> annexin A2 <strong>and</strong> S100A10. AIIt stimulated autoproteolysis<br />

<strong>of</strong> the plasmin Lys468-Gly469 bond <strong>and</strong> also catalyzed the reduction <strong>of</strong> the plasmin<br />

Cys462Cys541disulfide. These two reactions resulted in the release <strong>of</strong> angiostatin (Lys78-Lys468)<br />

from plasmin. However, oxidation <strong>of</strong> annexin A2 was not observed after knockdown <strong>of</strong> S100A10.<br />

Thus, AIIt acts as a plasmin reductase, catalyzing the release <strong>of</strong> angiostatin from plasmin. Results<br />

form these studies have established that S100A10 is a key regulatory protein <strong>of</strong> cell surface<br />

plasmin <strong>and</strong> angiostatin generation. Funded by the CIHR <strong>and</strong> NCIC.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 69


i058i<br />

<strong>Plasminogen</strong> Activator Inhibitor Type 2 Binds to S100A10 in<br />

Annexin 2 Heterotetramer <strong>and</strong> Prevents Annexin 2-dependent<br />

Plasmin In-site Formation by Inhibiting tPA<br />

Lobov S*, Croucher D, Ranson M<br />

School <strong>of</strong> Biological Sciences, University <strong>of</strong> Wollongong, NSW, Australia<br />

Presenting author e-mail: sergei@uow.edu.au<br />

<strong>Plasminogen</strong> Activator Inhibitor type 2 is a well known member <strong>of</strong> the serpin superfamily.<br />

Extracellular PAI-2 efficiently inhibits uPA <strong>and</strong> tPA in solution but there seem to be no<br />

physiological role for PAI-2 as tPA inhibitor. Interestingly however, PAI-2 as well as tPA <strong>and</strong><br />

plasminogen are capable <strong>of</strong> binding to annexin 2 heterotetramer (AIIt). AIIt acts as a receptor<br />

for tPA <strong>and</strong> plasminogen on the cell surface thereby promoting tPA-dependent plasmin in-site<br />

formation. The fact that PAI-2 is capable <strong>of</strong> binding to AIIt is known for more then a decade.<br />

However, neither the mechanism nor the physiological outcomes <strong>of</strong> the binding are known.<br />

Here we show that PAI-2 binds to AIIt via the S100A10 subunit <strong>and</strong> that the CD-loop <strong>of</strong> PAI-2<br />

plays a minor role in the binding. Furthermore, the interaction <strong>of</strong> PAI-2 with AIIt, unlike tPA <strong>and</strong><br />

plasminogen, is AIIt C-terminal lysine independent, which might involve novel binding sites in<br />

S100A10 <strong>and</strong> AIIt. In spite <strong>of</strong> seemingly different sites <strong>of</strong> binding to AIIt, PAI-2 <strong>and</strong> plasminogen<br />

could not simultaneously bind to AIIt. Furthermore, simultaneous binding <strong>of</strong> PAI-2 <strong>and</strong> tPA<br />

seemed to be able to exist only for the PAI-2/tPA complex. Enzyme kinetic analyses confirmed<br />

that PAI-2 inhibits AIIt-bound tPA in vitro <strong>and</strong> thus prevents plasmin in-site formation. We<br />

conclude that there is a complex relationship between PAI-2/plasminogen/tPA when AIIt is<br />

present. We propose that extracellular PAI-2 may have an important physiological role as a tPA<br />

inhibitor on the cell surface.<br />

70 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i059i<br />

Underst<strong>and</strong>ing the Structural Basis <strong>of</strong> the Differential-Receptor-<br />

Mediated Endocytosis Mechanisms <strong>of</strong> PAI-1 <strong>and</strong> PAI-2 in Cancer<br />

Cochran BJ* 1 , Lobov S 1 , Croucher D 2 , Ranson M 1<br />

1 Cancer Research Laboratory, School <strong>of</strong> Biological Sciences, University <strong>of</strong> Wollongong, Australia;<br />

2 Cancer Research Program, Garvan Institute <strong>of</strong> Medical Research, Australia<br />

Presenting author e-mail: blake@uow.edu.au<br />

The urokinase-type plasminogen activator (uPA) <strong>and</strong> its inhibitors, plasminogen activator<br />

inhibitor type-1 (PAI-1) <strong>and</strong> type-2 (PAI-2) play an important role in the progression <strong>of</strong> a number<br />

<strong>of</strong> forms <strong>of</strong> cancer. Whilst both PAI-1 <strong>and</strong> PAI-2 clear uPA activity from the cell surface through<br />

irreversible complex formation, uPA:PAI-1 initiates cell signalling events most likely associated<br />

with poor cancer prognosis. This activation is dependent on interactions between PAI-1 <strong>and</strong><br />

the very low density lipoprotein receptor (VLDLr). Upon internalisation, uPA:PAI-2 does not<br />

activate these cell signalling pathways, possibly accounting for the association between PAI-2<br />

expression <strong>and</strong> good prognosis in some cancer types. The mechanisms underlying this differential<br />

functionality are poorly understood <strong>and</strong> is believed to be due to the absence <strong>of</strong> LDLR binding<br />

sites in the PAI-2 moiety <strong>of</strong> the uPA:PAI-2 complex.<br />

This study aims to reconstitute the LDLR binding ability <strong>of</strong> PAI-1 in PAI-2 via the replacement<br />

<strong>of</strong> structural elements <strong>and</strong> residues <strong>of</strong> PAI-2 with those corresponding to PAI-1. To date, we<br />

have constructed an ahelix D domain swap which retained uPA inhibitory activity. We intend<br />

to compare the LDLR binding characteristics <strong>of</strong> this protein chimera to PAI-1. The findings <strong>of</strong><br />

this research may help explain the underlying reasons for the distinct biological roles <strong>of</strong> PAI-1<br />

<strong>and</strong> PAI-2 in cancer. A more complete underst<strong>and</strong>ing <strong>of</strong> the biological roles <strong>of</strong> PAI-2 will provide<br />

important information relating to the function <strong>of</strong> the plasminogen activation system in cancer <strong>and</strong><br />

may aid in the continuing development <strong>of</strong> highly specific <strong>and</strong> effective anti-uPA treatments.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 71


i060i<br />

A Low-glycemic-index Diet Reduces Plasma PAI-1 Activity<br />

in Overweight Women<br />

Jensen L* 1 , Krog-Mikkelsen I 2 , Sloth B 2 , Flint A 2 , Astrup A 2 , Raben A 2 , Tholstrup T 2 , Brünner N 1<br />

1 Department <strong>of</strong> Veterinary Pathobiology, 2 Department <strong>of</strong> Human Nutrition, Centre for Advanced Food<br />

Studies, Faculty <strong>of</strong> Life Sciences, University <strong>of</strong> Copenhagen, Frederiksberg, Denmark<br />

Presenting author e-mail: lje@life.ku.dk<br />

An elevated level <strong>of</strong> plasminogen activator inhibitor-1 (PAI-1) in plasma is a core feature <strong>of</strong> the<br />

metabolic syndrome. PAI-1 has been shown to decrease during weight loss. However, effects <strong>of</strong><br />

healthy diets on PAI-1 levels may not solely depend on weight loss. The relevance <strong>of</strong> glycemic<br />

index in preventing the metabolic syndrome is controversial.<br />

The purpose <strong>of</strong> the present study was to investigate the effect <strong>of</strong> 10 weeks intake <strong>of</strong> a low<br />

glycemic index (LGI) vs. a high glycemic index (HGI) high-carbohydrate, low fat ad libitum diet<br />

on plasma PAI-1 activity <strong>and</strong> antigen levels in overweight women.<br />

44 healthy overweight women (BMI 27.5 ± 0.2 kg/m2) were r<strong>and</strong>omly assigned to a parallel 10<br />

week intervention with a LGI (n=22) or HGI (n=22) diet. To study the postpr<strong>and</strong>ial effect <strong>of</strong> LGI<br />

vs. HGI diets a subgroup <strong>of</strong> 29 subjects (LGI, n=14; HGI n=15) were assigned to participate in an<br />

additional 4-h meal test on the last day <strong>of</strong> the intervention. Fasting blood samples were obtained<br />

before <strong>and</strong> after the 10 weeks, <strong>and</strong> postpr<strong>and</strong>ially at 10 weeks.<br />

PAI-1 activity in plasma decreased (P = 0.02) after the LGI diet, a decrease that was significantly<br />

different (P < 0.01) from the change after a HGI diet. Changes in PAI-1 antigen levels were<br />

not significantly different between any groups. Postpr<strong>and</strong>ial PAI-1 concentrations showed no<br />

significant differences between groups.<br />

A LGI diet reduces plasma PAI-1 activity, suggesting that a beneficial role <strong>of</strong> a LGI diet in<br />

overweight adults could be related to PAI-1.<br />

72 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i061i<br />

Urokinase Receptor-independent Signalling <strong>of</strong> the Urokinase-type<br />

<strong>Plasminogen</strong> Activator via Phosporylation <strong>of</strong> STAT1<br />

Ehart M* <strong>and</strong> Binder BR<br />

Department <strong>of</strong> Vascular <strong>Biology</strong> <strong>and</strong> Thrombosis Research, Medical University <strong>of</strong> Vienna, Austria<br />

Presenting author e-mail: monika.ehart@meduniwien.ac.at<br />

Phosphorylation <strong>of</strong> STAT-proteins is one possible way <strong>of</strong> signalling induced by the urokinase-<br />

type plasminogen activator (uPA). We analyzed the dependency <strong>of</strong> STAT1 phosphorylation on the<br />

presence <strong>of</strong> the urokinase receptor in mouse skin fibroblasts. We found that human uPA, which<br />

is assumed not to bind to the mouse urokinase receptor, induces STAT1 phosphorylation. Much<br />

to our surprise this phosphorylation was induced only in cells derived from urokinase receptor<br />

knock out mice, but not in cells derived from wild type mice. Since we <strong>and</strong> others have shown<br />

that gp130 is involved in uPA induced STAT signalling, we analyzed association <strong>of</strong> uPA with<br />

gp130. We could precipitate gp130 together with uPA from cells derived from urokinase receptor<br />

knock out mice but not from cells derived from wild type mice. We further analyzed urokinase<br />

receptor independent uPA signalling in the human kidney epithelial cell line HEK-293 <strong>and</strong> the<br />

human adenocarcinoma cell line MCF-7 <strong>and</strong> obtained data supporting the notion that gp130<br />

transuces uPA signals only in the absence <strong>of</strong> urokinase receptor.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 73


i062i<br />

Urokinase Receptor Promotes Neo-Angiogenesis through<br />

its Ser88-Arg-Ser-Arg-Tyr92 Chemotactic Sequence<br />

Longanesi-Cattani I* 1 , Bifulco K 1 , Cantelmo AR 1 , Di Carluccio G 1 , Spina R 1 , Liguori E 1 , Stoppelli MP 2 ,<br />

Carriero MV 1<br />

1 Department <strong>of</strong> Experimental Oncology, National Cancer Institute <strong>of</strong> Naples, Italy;<br />

2 Institute <strong>of</strong> Genetics <strong>and</strong> Biophysics ‘’Adriano Buzzati-Traverso,’ Naples, Italy<br />

*Presenting author e-mail: immalonganesi@libero.it<br />

Angiogenesis is a highly coordinated process required for normal development, in response<br />

to injury <strong>and</strong> for tumor growth. This process is sustained by a tightly regulated motility <strong>of</strong><br />

endothelial cells. Angiogenesis is regulated by chemotactic stimuli <strong>and</strong> requires the activation <strong>of</strong><br />

several signalling pathways that converge on cytoskeletal remodelling. The Ser88-Arg-Ser-Arg-<br />

Tyr92 chemotactic sequence <strong>of</strong> the urokinase receptor binds to formyl peptide receptor-like-1<br />

(FPRL-1), a G protein-coupled cell receptor (Resnati et al. 2002). Furthermore, we have found<br />

that the Ser-Arg-Ser-Arg-Tyr (SRSRY) peptide binds to the high affinity formyl peptide receptor<br />

(FPR), <strong>and</strong> specifically promotes cytoskeletal rearrangements <strong>and</strong> directional cell migration in<br />

a vitronectin receptor-dependent manner (Gargiulo et al. 2005). Since Human Umbelical Vein<br />

Endothelial Cells (HUVEC) express FPRL-1 as well as avb3 vitronectin receptors, we investigated<br />

on the possibility that SRSRY may triggers neo-angiogenesis. By conventional Boyden chamber<br />

assays, we found that SRSRY promotes directional cell migration <strong>of</strong> HUVEC in a dose-dependent<br />

manner, with an extent similar to VEGF. Maximal chemotactic effect is reached at 10 nM<br />

SRSRY concentration. In a tube-formation assay, SRSRY shows a remarkable ability to promote<br />

angiogenesis <strong>of</strong> endothelial cells. According to cell migration data, pro-angiogenic effect is dosedependent<br />

<strong>and</strong> peaks at 10 nM SRSRY. Both endothelial cell migration <strong>and</strong> in vitro angiogenesis<br />

promoted by SRSRY are inhibited by cell pre-incubation with blocking anti-avb3 monoclonal<br />

antibodies, indicating that SRSRY exhibits an avb3-dependent pro-angiogenetic effect. Although<br />

preliminary, these data suggest the development <strong>of</strong> SRSRY-derived pharmacological compounds<br />

to be employed in tissue repairing <strong>and</strong> cardiovascular diseases.<br />

74 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i063i<br />

The Density Enhanced Phosphatase 1 (DEP-1) Down-Modulates<br />

Urokinase Receptor (uPAR) Surface Expression in Confluent<br />

Endothelial Cells<br />

Brunner PM* 1 , Heier PC 1 , Prager GW 1,2 , Mihaly J 1 , Priglinger U 3 , Binder BR 1<br />

1Department <strong>of</strong> Vascular <strong>Biology</strong> <strong>and</strong> Thrombosis Research, Center for Biomolecular Medicine <strong>and</strong><br />

Pharmacology, Medical University Vienna, Austria;<br />

2 Clinical Division <strong>of</strong> Oncology, Department <strong>of</strong> Medicine I, Medical University Vienna, Austria;<br />

3 Department <strong>of</strong> Emergency Medicine, Medical University Vienna, Austria<br />

Presenting author e-mail: patricia.heier@univie.ac.at<br />

It is known that a proteolytic machinery is indispensable for ordered endothelial cell migration<br />

<strong>and</strong> tissue invasion. Such machinery is largely provided by the prourokinase-urokinase-plasmin<br />

system. Endothelial cells (ECs) in vessels are forming a confluent layer that can be mimicked<br />

by confluent ECs in cell culture. Compared to sparse cultures we could show that the surface<br />

expression <strong>of</strong> uPAR is decreased by 40%. To examine the cause <strong>of</strong> this decrease we investigated<br />

the MAP-Kinase-pathway being the major regulatory system for uPAR-expression.<br />

We found that between sparse (50,000 cells/cm2), subconfluent (100,000 cells/cm2) <strong>and</strong> confluent<br />

cells (150,000 cells/cm2) there was no change in ERK1/2 expression or its phosphorylation.<br />

Therefore we analyzed phosphatases as possible c<strong>and</strong>idates for density dependent uPARexpression.<br />

While Dual-Specificity Phosphatase 1 (DUSP1) was not <strong>and</strong> DUSP4 was only<br />

borderline upregulated with density, Density Enhanced Phosphatase 1 (DEP-1) significantly<br />

increased with cell density. Transfection <strong>of</strong> ECs with expression plasmids for wild-type<br />

<strong>and</strong> mutated forms <strong>of</strong> DEP-1 revealed that wild-type DEP-1 but not forms mutated in the<br />

phosphatase-domain decreased uPAR-mRNA <strong>and</strong> cell surface expression. To analyse in detail<br />

how DEP-1 interferes with uPAR-expression we applied reporter assays using plasmids encoding<br />

ERK1 <strong>and</strong> different DEP-1 forms. We found that DEP-1 influenced the MAP-Kinase pathway<br />

downstream <strong>of</strong> ERK1/2 in addition to its tyrosine kinase dephosphorylating activity. From this<br />

data we conclude that the quiescent state <strong>of</strong> contact inhibited ECs is at least partially mediated by<br />

DEP-1 (<strong>and</strong> possibly also by DUSP4) that inhibits the MAP-Kinase pathway also downstream <strong>of</strong><br />

ERK1/2.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 75


i064i<br />

Detection <strong>and</strong> Prevention <strong>of</strong> Hepatic Fibrosis Targeting<br />

Proteolytic TGF-b <strong>Activation</strong> Reaction<br />

Kojima S*<br />

<strong>Molecular</strong> <strong>Cellular</strong> Pathology Research Unit, RIKEN, Japan<br />

Presenting author e-mail: skojima@postman.riken.go.jp<br />

Transforming growth factor (TGF)-b, the most fibrogenic cytokine participating in the<br />

pathogenesis <strong>of</strong> liver diseases, is produced as a high molecular weight latent form, <strong>and</strong> thus must<br />

be activated before exerting its biological activities. TGF-b activation is the reaction, by which 25<br />

kD active TGF-b molecule is released from the latent complex. We showed that TGF-b is activated<br />

by plasmin (PLN) <strong>and</strong> plasma kallikrein (PLK) during pathogenesis <strong>of</strong> liver fibrosis <strong>and</strong> impaired<br />

liver regeneration, respectively, <strong>and</strong> that blockage <strong>of</strong> these activation reactions with low molecular<br />

weight protease inhibitors prevented the development <strong>of</strong> the diseases in animal models.<br />

PLN <strong>and</strong> PLK cleaved between K56-L57 <strong>and</strong> R58-L59 in LAP portion <strong>of</strong> human latent TGF-b1,<br />

respectively. We produced antibodies that specifically recognize the neo-epitopes formed by<br />

protease degradation, namely the cut ends <strong>of</strong> each cleavage site. The anti-R58 antibodies strongly<br />

stained liver sections from patients with fulminant hepatitis compared to normal liver sections<br />

from patients that died from pulmonary embolism. On the other h<strong>and</strong>, the anti-L59 antibodies<br />

were successively used to establish ELISA to detect LAP degradation fragments in mouse serum.<br />

These results suggest a key role for PLN/PLK in the generation <strong>of</strong> active TGF-b, namely that a<br />

PLN/PLK-dependent activation reaction occurs around hepatic stellate cells to produce active<br />

TGF-b, which may induce fibrosis <strong>and</strong> inhibit the proliferation <strong>of</strong> hepatocytes, thereby addressing<br />

a potential use for PLN/PLK inhibitors in hepatitis therapy. These results also suggest a potential<br />

usage <strong>of</strong> a LAP degradate as a biomarker for liver fibrosis.<br />

76 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i065i<br />

Domain 1 <strong>of</strong> uPAR Is Required for its Morphological<br />

<strong>and</strong> Functional b2 Integrin-mediated Connection with<br />

Actin Cytoskeleton in Human Endothelial Cells<br />

Del Rosso M* 1 , Fibbi G 1 , Margheri F 1 , Serratì S 1 , Pucci M 1 , Manetti M 2 , Ibba-Manneschi L 2<br />

1 Department <strong>of</strong> Experimental Pathology <strong>and</strong> Oncology, University <strong>of</strong> Florence, Italy;<br />

2 Department <strong>of</strong> Anatomy, Histology <strong>and</strong> Forensic Medicine, University <strong>of</strong> Florence, Italy<br />

Presenting author e-mail: delrosso@unifi.it<br />

We have previously shown that MMP12-dependent cleavage <strong>of</strong> uPAR domain 1 blocks<br />

angiogenesis in Systemic Sclerosis (SSc) endothelial cell. Since integrin association accounts<br />

for uPAR invasion required in angiogenesis, this study was undertaken to show whether fullsize<br />

<strong>and</strong> truncated uPAR differentially associated to integrins <strong>and</strong> with motor components<br />

<strong>of</strong> the cytoskeleton. Truncated <strong>and</strong> full-size uPAR <strong>and</strong> its association with integrins <strong>and</strong> actin<br />

cytoskeleton in SSc <strong>and</strong> normal (N) microvascular endothelial cells (MVEC), isolated from skin<br />

biopsies, were studied by Confocal Laser Scanning Microscopy <strong>and</strong> immuno-precipitation.<br />

Integrin composition <strong>of</strong> endothelial cells was studied by RT-PCR. Cell migration <strong>and</strong> capillary<br />

morphogenesis were studied on fibrin(ogen) substrates. Involvement <strong>of</strong> Rac <strong>and</strong> Cdc42 was<br />

showed by Western blot.<br />

Only full-size uPAR is connected with actin cytoskeleton in endothelial cells. Such connection<br />

is mediated by uPAR-associated alphaM-beta2 <strong>and</strong> alphaX-beta2 integrins <strong>and</strong> is absent in<br />

endothelial cells <strong>of</strong> SSc patients. The cleaved receptor does not associate with beta2 integrins<br />

<strong>and</strong> with actin. Beta3 integrins associated with both full-size <strong>and</strong> cleaved uPAR at focal contacts.<br />

uPAR-beta2 integrins uncoupling in N-MVEC impaired activation <strong>of</strong> Rac <strong>and</strong> Cdc42, which<br />

mediate uPAR-dependent cytoskeletal rearrangements <strong>and</strong> cell motility, as well as integrin<br />

engagement-delivered signals to actin cytoskeleton. Invasion <strong>and</strong> capillary morphogenesis on<br />

fibrin(ogen)-coated substrates indicated that uPAR ligation by uPA empowers beta2 <strong>and</strong> beta3<br />

integrin-dependent fibrin(ogen) invasion <strong>and</strong> that such system is impaired in SSc endothelial cell.<br />

Conclusions. uPAR truncation <strong>and</strong> the following loss <strong>of</strong> beta2 integrin-mediated uPAR connection<br />

with actin cytoskeleton account for reduced angiogenic properties <strong>of</strong> SSc endothelial cells.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 77


i066i<br />

The uPA/uPAR/Vn Pathway <strong>of</strong> Signaling to MAPK-activation<br />

Sarra Ferraris GM* 1 , Madsen C 1 , Sidenius N 1,2<br />

1 IFOM Foundation, FIRC Institute <strong>of</strong> <strong>Molecular</strong> Oncology, Milan Italy;<br />

2 <strong>Molecular</strong> Genetics Unit, DIBIT, Università Vita-Salute San Raffaele, Milan Italy<br />

Presenting author e-mail: gianmaria.sarraferraris@ifom-ieo-campus.it<br />

The urokinase-type plasminogen activator receptor (uPA(R)), despite the lack <strong>of</strong> a transmembrane<br />

domain, behaves as a signaling molecule resulting in increased ERK1/2-activation when<br />

overexpressed or upon uPA-binding. In accordance, we find that CHO <strong>and</strong> 293 cells overexpressing<br />

uPAR display increased basal ERK1/2-activation.<br />

Two different alanine substitution in uPAR (W32A <strong>and</strong> T54A), which impair Vn-binding, both<br />

failed to cause increased basal ERK1/2-activation. The addition <strong>of</strong> uPA to cells expressing the<br />

T54A receptor induced rapid binding to vitronectin <strong>and</strong> ERK1/2 activation, while uPA binding<br />

to cells expressing the W32A mutant failed to induce both vitronectin binding <strong>and</strong> ERK1/2<br />

activation. Both mutants bind uPA equally well <strong>and</strong> these data thus allow us to conclude<br />

that uPA-induced ERK1/2 activation is mediated by the induction <strong>of</strong> the uPAR/vitronectininteraction.<br />

Using cells expressing theT54A receptor we compared the kinetics <strong>of</strong> uPAR/Vn-induced ERK1/2<br />

activation with the ones triggered by other signalling receptors.<br />

uPA <strong>and</strong> LPA treatments caused a sustained ERK1/2 activation while EGF an Mn2+ displayed a<br />

peek at 5 minutes followed by a rapid decrease <strong>of</strong> phosphorylated ERK1/2 levels.<br />

Moreover EGF receptor inhibitors, LPA1 receptor inhibitor <strong>and</strong> Pertussis Toxin did not affect uPAuPAR<br />

signalling pathway while PI3 kinase inhibitors decreased partially but significantly ERK1/2<br />

activation in response to uPA <strong>and</strong> LPA.<br />

These data underline the importance <strong>of</strong> Vn binding in the uPA-uPAR signalling, suggesting a new<br />

way, which is independent <strong>of</strong> the EGFR <strong>and</strong> pertussis toxin sensitive GPCR’s, by which uPAR<br />

may signal to ERK1/2.<br />

78 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i067i<br />

Regulation <strong>of</strong> tPA <strong>and</strong> PAI-1 Gene Expression in Astrocytes<br />

Hultman K* 1 , Tjärnlund-Wolf A 1 , Blomstr<strong>and</strong> F 1 , Nilsson M 1 , Medcalf R 2 , Jern C 1<br />

1 Institute <strong>of</strong> Neuroscience <strong>and</strong> Physiology, the Sahlgrenska Academy at Göteborg University, Sweden;<br />

2 Australian Centre for Blood Diseases, Monash University, Alfred Medical Research <strong>and</strong> Education<br />

Precinct (AMREP), Australia<br />

Presenting author e-mail: karin.hultman@neuro.gu.se<br />

We have shown that PKC activation acts in synergy with retinoic acid (RA) to induce tissue-type<br />

plasminogen activator (tPA) gene expression in human endothelial cells. The aim <strong>of</strong> the present<br />

study was to investigate whether tPA is regulated in a similar manner in astrocytes, as well as to<br />

characterize plasminogen activator inhibitor type 1 (PAI-1) gene expression in these cells.<br />

Native human astrocytes were treated with RA, PKC activator (PMA), PKA activator (forskolin),<br />

cytokines or growth factors for 3, 6, 14 <strong>and</strong> 20 hours. RA or PMA alone induced a 3-fold upregulation<br />

<strong>of</strong> tPA at 20 hours. Combined treatment resulted in a 9-fold induction <strong>and</strong> increased<br />

intracellular storage pools <strong>of</strong> tPA as visualized by immunocytochemistry. Pretreatment with<br />

actinomycin D or cycloheximide completely blocked both RA <strong>and</strong> PMA mediated tPA mRNA<br />

increase. Cytokines <strong>and</strong> growth factors caused a general weak up-regulation <strong>of</strong> tPA <strong>and</strong> PAI-1 at<br />

3 <strong>and</strong> 6 hours followed by a slight down-regulation at 20 hours. Forskolin induced a strong timedependent<br />

down-regulation <strong>of</strong> both proteins. For all treatments, similar inductions were observed<br />

at the mRNA (RT-PCR) <strong>and</strong> protein level (ELISA).<br />

We show for the first time that in astrocytes, RA <strong>and</strong> PKC activation induce a strong up-regulation<br />

<strong>of</strong> tPA gene expression <strong>and</strong> that this response is dependent on both gene transcription <strong>and</strong> de<br />

novo protein synthesis. As we <strong>and</strong> others have identified functional variants in the tPA <strong>and</strong> PAI-1<br />

promoters, we will now proceed by investigating whether there is an allele-specific regulation in<br />

human astrocytes.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 79


i068i<br />

Identification <strong>of</strong> a Mitotic Epitope in the Domain 2 <strong>of</strong> the<br />

Urokinase Receptor (uPAR)<br />

Degryse B* 1 , Eden G 2 , Arnaudova R 1 , Furlan F 1 , Blasi F 1,2<br />

1Department <strong>of</strong> <strong>Molecular</strong> <strong>Biology</strong> <strong>and</strong> Functional Genomics, DIBIT, University Vita Salute San<br />

Raffaele, Milan, Italy;<br />

2 IFOM, FIRC Institute <strong>of</strong> <strong>Molecular</strong> Oncology, Milan, Italy<br />

Presenting author e-mail: degryse.bernard@hsr.it<br />

The first characterized chemotactic sequence <strong>of</strong> uPAR (SRSRY) is located in the linker region<br />

between domain 1 <strong>and</strong> 2. SRSRY peptide promotes cell migration through binding to receptors<br />

such as FPRL1. Recently, we identified in the domain 2 another chemotactic epitope that we<br />

named D2A, 130IQEGEEGRPKDDR142 (Degryse et al., 2005, J. Biol. Chem. 280, 24792-24803). D2A<br />

synthetic peptide dose-dependently stimulates cell migration with a maximum at 1 pM. We have<br />

also shown that D2A promotes cell migration through binding to avb3 <strong>and</strong> a5b1 integrins <strong>and</strong><br />

activation <strong>of</strong> integrin-dependent signaling.<br />

Since uPAR can regulate cell proliferation, we have started investigating the effects <strong>of</strong> D2A on<br />

cell growth. D2A stimulates cell growth in a dose-dependent manner in various cell types <strong>and</strong><br />

species, whereas a scrambled version <strong>of</strong> D2A has no effect. In rat smooth muscle cells, the optimal<br />

dose is 1-10 pM while in human HT29 <strong>and</strong> HT1080 cells it is 100 pM. These results demonstrate<br />

that D2A is a mitogen at doses that are slightly higher compared to the optimal dose required for<br />

chemotaxis.<br />

We have identified a mitotic/chemotactic epitope in the domain 2 <strong>of</strong> uPAR showing that this<br />

region is particularly important for uPAR-induced signaling. These data suggest that uPAR is a<br />

membrane-bound mitogen acting through binding to other membrane receptor(s).<br />

80 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i069i<br />

Vitronectin Inhibits <strong>Plasminogen</strong> Activator Inhibitor-1<br />

(PAI-1)-Induced Chemotaxis by Blocking PAI-1<br />

Binding to the LDL Receptor-Related Protein<br />

Neels JG 1,2 , Kamikubo Y 1,3 , Degryse B* 1,4<br />

1Division <strong>of</strong> Vascular <strong>Biology</strong>, Department <strong>of</strong> Cell <strong>Biology</strong>, The Scripps Research Institute, La Jolla,<br />

California, USA;<br />

Present addresses:<br />

2 Department <strong>of</strong> Medicine, University <strong>of</strong> California San Diego, La Jolla, California, USA;<br />

3Department <strong>of</strong> <strong>Molecular</strong> <strong>and</strong> Experimental Medicine, The Scripps Research Institute, La Jolla,<br />

California, USA;<br />

4Department <strong>of</strong> <strong>Molecular</strong> <strong>Biology</strong> <strong>and</strong> Functional Genomics, DIBIT, University Vita-Salute San<br />

Raffaele, Milan, Italy<br />

Presenting author e-mail: degryse.bernard@hsr.it<br />

We have previously shown (Degryse et al., 2004, J. Biol. Chem. 279, 22595-22604) that PAI-1<br />

activates the Jak/Stat signaling pathway <strong>and</strong> stimulates cell migration by binding to the LDL<br />

receptor-related protein (LRP), a large endocytic receptor. Since vitronectin (VN) influences most<br />

<strong>of</strong> the biological functions <strong>of</strong> PAI-1, we explored the effects <strong>of</strong> VN on PAI-1-induced signaling <strong>and</strong><br />

cell migration. In fact, VN inhibits PAI-1-promoted signaling <strong>and</strong> chemotaxis. VN acts by binding<br />

to PAI-1 inhibiting the activation <strong>of</strong> the Jak/Stat pathway <strong>and</strong> cell migration. VN exerts these<br />

inhibitory effects by blocking PAI-1 binding to LRP, its motogenic receptor.<br />

We have unveiled a new inhibitory role <strong>of</strong> VN which has all the hallmarks <strong>of</strong> a molecule that can<br />

regulate every steps <strong>of</strong> the cell migration cycle by providing ‘’Stop’’ (under the form <strong>of</strong> VN/PAI-<br />

1 complex) <strong>and</strong> ‘’Go’’ signals to the cell. VN shares striking similarities with urokinase, which is<br />

the other main lig<strong>and</strong> <strong>of</strong> uPAR <strong>and</strong> can also generate ‘’Stop’’ <strong>and</strong> ‘’Go’’ signals. Furthermore, it is<br />

intriguing that urokinase <strong>and</strong> VN induce ‘’Stop’’ signals by creating uPA/PAI-1 <strong>and</strong> VN/PAI-1<br />

complexes that block uPAR-, integrin- <strong>and</strong> LRP-dependent cell migration.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 81


i070i<br />

Estradiol Inhibits EGF-induced Cell Migration <strong>and</strong> uPAR Expression in<br />

Estrogen Receptor-a Negative, GPR30 Positive Ovarian Cancer Cells<br />

Henic E* 1 , Noskova V 1 , Høyer-Hansen G 2 , Hansson S 1 , Casslén B 1<br />

1 Department <strong>of</strong> Gynecology & Obstetrics, University Hospital, Lund, Sweden;<br />

2 Finsen Laboratory, Rigshospitalet, Copenhagen, Denmark<br />

Presenting author e-mail: emir.henic@med.lu.se<br />

EGF is a stimulator <strong>of</strong> proliferation as well as migration in ovarian cancer cells. In contrast, the<br />

effects <strong>of</strong> estrogen are incompletely explored <strong>and</strong> some results are contradictory. In addition to<br />

genomic effects via classical nuclear estrogen receptors (ER), accumulating evidence suggest that<br />

estrogens initiate rapid non-genomic signaling through cell membrane receptors. GPR30 is a<br />

newly identified receptor with all characteristics <strong>of</strong> a membrane ER.<br />

In this study we show that estradiol attenuates the stimulatory effect <strong>of</strong> EGF on both cell<br />

migration <strong>and</strong> uPAR expression in ovarian cancer cell lines. We have previously shown that EGF<br />

up-regulates uPAR expression via three distinct mechanisms. Studying each <strong>of</strong> these mechanisms<br />

in the presence <strong>of</strong> estradiol, we found no inhibition <strong>of</strong> neither the production <strong>of</strong> uPAR, i.e.<br />

the level <strong>of</strong> uPAR mRNA, nor the elimination <strong>of</strong> cell surface uPAR, i.e. by internalization for<br />

lysosomal degradation or by shedding to the medium. Instead estradiol inhibited the very rapid<br />

increase <strong>of</strong> detergent extractable uPAR, which occurs within minutes <strong>of</strong> EGF stimulation, <strong>and</strong> is<br />

likely to represent mobilization <strong>of</strong> uPAR from detergent resistant domains, like lipid rafts. ICI<br />

182,780, which is a nuclear ER antagonist, has agonistic properties with GPR30. This compound,<br />

like estradiol, inhibited the immediate effect <strong>of</strong> EGF on uPAR expression, <strong>and</strong> also did not inhibit<br />

that effect <strong>of</strong> estradiol. OVCAR-3 cells are ERa negative, but express mRNA for GPR30.<br />

Thus, our observations suggest that modulation <strong>of</strong> the EGF response by estradiol is mediated by<br />

GPR30.<br />

82 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i071i<br />

Methylation <strong>of</strong> the PAI-1 Gene in Oral Squamous<br />

Cell Carcinomas <strong>and</strong> Normal Oral Mucosa<br />

Gao S 1 , Krogdahl A 2 , Sørensen JA 3 , Dabelsteen E 4 , Andreasen PA* 1<br />

1 Department <strong>of</strong> <strong>Molecular</strong> <strong>Biology</strong>, University <strong>of</strong> Aarhus, Aarhus, Denmark;<br />

2 Department <strong>of</strong> Pathology, Odense University Hospital, Odense, Denmark;<br />

3 Department <strong>of</strong> Plastic Surgery, Odense University Hospital, Odense, Denmark;<br />

4 School <strong>of</strong> Dentistry, University <strong>of</strong> Copenhagen, Copenhagen, Denmark<br />

Presenting author e-mail: pa@mb.au.dk<br />

It has become clear that CpG methylation plays an important role in development <strong>and</strong><br />

progression <strong>of</strong> cancer <strong>and</strong> the associated changes in gene expression. We have become interested<br />

in the possibility that decreased CpG methylation <strong>of</strong> specific genes may lead to over-expression<br />

in tumours <strong>and</strong> have studied the plasminogen activator inhibitor-1 (PAI-1) gene in this respect.<br />

The PAI-1 gene has 25 CpG sites within 960 bp around the transcription initiation site. We studied<br />

CpG methylation <strong>of</strong> the PAI-1 gene by bisulfite sequencing <strong>and</strong> PAI-1 mRNA levels by real-time<br />

RT-PCR. In a series <strong>of</strong> 20 cases <strong>of</strong> oral carcinomas <strong>and</strong> matched samples <strong>of</strong> adjacent histologically<br />

normal tissue from the same patients, we observed that the methylation frequency in tumours<br />

was lower than in the adjacent non-tumour tissue. By real time RT-PCR analysis, 17 <strong>of</strong> 20 patients<br />

with oral carcinoma were found to have between 2.5 <strong>and</strong> 50 fold increased PAI-1 mRNA levels,<br />

as compared with the adjacent non-tumour tissue. The mRNA level in the tumours was inversely<br />

correlated with decreased methylation frequency in the tumours. We conclude that demethylation<br />

<strong>of</strong> the PAI-1 gene promoter region may contribute to the higher expression <strong>of</strong> PAI-1 in the tumour<br />

than in surrounding histologically normal tissue in some oral carcinoma patients. The methylation<br />

at some specific CpG sites may play an important role in the regulation <strong>of</strong> PAI-1 gene expression.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 83


i072i<br />

Urokinase Signaling through Its Receptor Promotes<br />

Invasiveness <strong>and</strong> Metastasis <strong>of</strong> Pancreatic Cancer Cells<br />

Xue A*, Xue M, Jackson C, Song E, Allen BJ, Smith RC<br />

The University <strong>of</strong> Sydney, Department <strong>of</strong> Surgery, Royal North Shore Hospital, NSW, Australia<br />

Presenting author e-mail: aiqunn@med.usyd.edu.au<br />

Urokinase-type plasminogen activator (uPA) system plays an important role in tumour<br />

pathogenesis. However, non-specific inhibition <strong>of</strong> all uPA system members causes side-effects,<br />

so it is necessary to clarify the specific function <strong>of</strong> uPA system members to provide the basis for<br />

selective blockade <strong>and</strong> reduction <strong>of</strong> therapeutic side-effects. In this study, we explored the role <strong>of</strong><br />

uPA, uPA receptor (uPAR) <strong>and</strong> PA inhibitor (PAI) in pancreatic ductal adenocarcinoma (PDAC)<br />

pathogenesis in vitro. Cox regression analysis showed that uPAR was independently associated<br />

with shorter overall patient survival, whereas increased levels <strong>of</strong> tumour-associated PAI-2 is a<br />

potential indicator <strong>of</strong> good outcome for patients. Receiver-operating characteristic area under the<br />

curve (ROC AUC) further confirmed that uPAR mRNA levels, as a signer factor, had achieved<br />

the highest AUC (AUC=0.84, p < 0.0001 ) in PDAC. When analysis <strong>of</strong> uPAR was combined with<br />

uPA, AUC was significantly increased up to 0.93, with sensitivity <strong>of</strong> 88.2% <strong>and</strong> specificity <strong>of</strong> 87.5%<br />

(p < 0.000). We also found that silencing uPAR by the specific small interfering RNA (siRNA)<br />

significantly inhibits cell proliferation, induces G0-G1 arrest <strong>and</strong> cell apoptosis, <strong>and</strong> suppresses<br />

cell migration in pancreatic cancer cells. Moreover, suppression <strong>of</strong> uPAR or uPA expression<br />

differentially regulates the activation <strong>of</strong> MAP kinases. However, specific blocking PAI-2 has not<br />

been observed significant impact on pancreatic cancer cell proliferation <strong>and</strong> migration. Taken<br />

together, our data indicate that uPAR may be important biomarker for human pancreatic cancer<br />

progression <strong>and</strong> prognosis as well as potentially selective therapeutic target.<br />

84 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i073i<br />

Tissue <strong>Plasminogen</strong> Activator Induces Cell Proliferation in Pancreatic<br />

Cancer by a Non-catalytic Mechanism that Requires ERK1/2 <strong>Activation</strong><br />

through Epidermal Growth Factor Receptor <strong>and</strong> Annexin A2<br />

Ortiz-Zapater E 1,2 , Peiró S 1,2 , Roda O 1,2 , Corominas JM 3 , Aguilar S 1 , Ampurdanés C 1 , Real FX 1,2 ,<br />

Navarro P* 1<br />

1 Unitat de Biologia Cel-lular i <strong>Molecular</strong>, IMIM-prbb, Barcelona, Spain;<br />

2 Departament de Ciències Experimentals i de la Salut, Facultat de Ciències de la Salut i de la Vida,<br />

Universitat Pompeu Fabra, Barcelona, Spain;<br />

3 Departament de Patologia, Hospital del Mar, UAB, Barcelona, Spain<br />

Presenting author e-mail: pnavarro@imim.es<br />

Pancreatic cancer is one <strong>of</strong> the most aggressive human tumors, being the fifth-leading cause <strong>of</strong><br />

cancer death in the developed countries. Our previous data have shown that tissue plasminogen<br />

activator (tPA) is overexpressed in pancreatic ductal adenocarcinomas, playing a critical role<br />

in several events associated to tumor progression such as cell proliferation, invasion <strong>and</strong><br />

angiogenesis. Prior work supports the notion that the effects <strong>of</strong> tPA on cell invasion require its<br />

proteolytic activity. Here, we identify the molecular mechanism responsible for the proliferative<br />

effects <strong>of</strong> tPA on pancreatic tumor cells. tPA activates the ERK1/2 signalling pathway in a<br />

manner that is independent <strong>of</strong> its catalytic activity. We also show that at least two membrane<br />

receptors, EGFR <strong>and</strong> AnxA2, which are overexpressed in pancreatic cancer, are involved in<br />

the transduction <strong>of</strong> tPA signalling in pancreatic tumors, suggesting the establishment <strong>of</strong> an<br />

amplification loop involved in tumor cell proliferation. These results add novel insights into the<br />

non-catalytic functions <strong>of</strong> tPA in cancer <strong>and</strong> the molecular mechanisms involved in its effects on<br />

cell proliferation.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 85


i074i<br />

Modulation <strong>of</strong> Lung Carcinoma Cell Lines <strong>and</strong> Primary<br />

Cultures Migration by uPA-Derived <strong>and</strong> EGFR Inhibitors<br />

Franco P* 1 , Mancini A 2 , Votta G 1 , Caputi M 2 , Stoppelli MP 1<br />

1 Institute <strong>of</strong> Genetics <strong>and</strong> Biophysics ‘’Adriano Buzzati-Traverso’’, National Research Council, Naples,<br />

Italy;<br />

2 Department <strong>of</strong> Cardiothoracic <strong>and</strong> Respiratory Sciences, Second University <strong>of</strong> Naples, Naples,Italy<br />

Presenting author e-mail: franco@igb.cnr.it<br />

The detachment <strong>of</strong> malignant cells from primary tumors <strong>and</strong> their subsequent migration to<br />

distant sites, including intravasation <strong>and</strong> extravasation, leads to tumor dissemination. The<br />

uPA/uPAR system plays a crucial role in tumor progression <strong>and</strong> invasion. Over-expression <strong>of</strong><br />

uPAR correlates with a poor prognosis in many neoplastic conditions, including lung tumors,<br />

suggesting that the specific inhibition <strong>of</strong> the uPA/uPAR system or/<strong>and</strong> the relative interactors<br />

may be a useful strategy to prevent metastasis. Among the functional partners <strong>of</strong> uPAR is the<br />

EGFR, which is over-expressed in NSCL (non-small-cell-lung) cancers <strong>and</strong> correlates with an high<br />

metastatic ability <strong>of</strong> the tumor. In particular, the EGFR is a widely recognised molecular target for<br />

anti-neoplastic therapies.<br />

In this study, we investigated the functional cross-talk between the two systems <strong>and</strong> the<br />

possibility to simultaneously inhibit both receptors by the aid <strong>of</strong> uPAR or EGFR inhibitors.<br />

In particular, we demostrated that both epidermoid carcinoma Calu-1 <strong>and</strong> squamous<br />

mucoepidermoid NCI-H-292 cell lines pre-treated with anti-EGFR antibodies, fail to migrate not<br />

only toward EGF but also toward uPA. Similarly, pre-treatment <strong>of</strong> the same cells with anti-uPAR<br />

antibodies, induce a strong inhibition <strong>of</strong> both uPA- <strong>and</strong> EGF-dependent migration.<br />

We also tested whether the anti-proliferative pharmacological inhibitors <strong>of</strong> EGFR kinase, namely<br />

IRESSA or gefitinib (ZD1839) <strong>and</strong> tyrphostin (AG-1478) may interfere with the Calu-1 <strong>and</strong> the<br />

NCI-H-292 cell migration. These inhibitors prevent not only EGF-dependent but also uPAdependent<br />

migration, suggesting that EGFR tyrosine kinase activity is required for uPA-induced<br />

directional migration.<br />

To further investigate the relationship between the two systems, we employed non-transformed<br />

human retinal pigment epithelial (RPE) <strong>and</strong> breast carcinoma MCF-7 cell lines in which uPAR<br />

expression level is stably reduced by the RNA-interference technology. In these clones, EGFR<br />

levels are not modified. However, in RPE <strong>and</strong> MCF-7 clones EGF-dependent migration is strongly<br />

impaired, confirming that the uPAR is required to the EGF/EGFR-dependent migration.<br />

These results show that the EGF/EGFR <strong>and</strong> uPA/uPAR systems are functionally interdependent<br />

in lung carcinoma cell migration <strong>and</strong> suggest that targeting either one may affect both systems.<br />

86 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i075i<br />

A Novel Role <strong>of</strong> Ku80 in Regulation <strong>of</strong> PAI-1 Gene Expression<br />

in Migrating Endothelial Cells Induced by Thymosin b4<br />

Bednarek R* 1 , Boncela J 1 , Smolarczyk K 1 , Cierniewski CS 1,2<br />

1 Center <strong>of</strong> Medical <strong>Biology</strong>, Polish Academy <strong>of</strong> Science, Lodz, Pol<strong>and</strong>;<br />

2 Department <strong>of</strong> Medical <strong>and</strong> <strong>Molecular</strong> Biophysics, Medical University, Lodz, Pol<strong>and</strong><br />

Presenting author e-mail: rbednarek@cbm.pan.pl<br />

Our data demonstrate that increased intracellular expression <strong>of</strong> thymosin b4 (Tb4) is<br />

necessary <strong>and</strong> sufficient to induce PAI-1 gene expression in endothelial cells. To describe the<br />

mechanism <strong>of</strong> this effect, we produced Tb4 mutants with impaired functional motifs <strong>and</strong> tested<br />

their intracellular location <strong>and</strong> activity. Cytoplasmic distributions <strong>of</strong> Tb4(AcSDKPT/4A),<br />

Tb4(KLKKTET/7A), <strong>and</strong> Tb4(K16A) mutants fused with GFP did not differ significantly<br />

from those <strong>of</strong> wild Tb4. Overexpression <strong>of</strong> Tb4, Tb4(AcSDKPT/4A) <strong>and</strong> Tb4(K16A) affected<br />

intracellular formation <strong>of</strong> actin filaments. As expected, Tb4(K16A) uptake by nuclei was impaired.<br />

On the other h<strong>and</strong>, overexpression <strong>of</strong> Tb4(KLKKTET/7A) resulted in developing the actin<br />

filament network typical <strong>of</strong> adhering cells, indicating that the mutant lacked the actin binding<br />

site. The mechanism by which intracellular Tb4 induced the PAI-1 gene did not depend upon<br />

the N-terminal tetrapeptide AcSDKP, <strong>and</strong> depended only partially on its ability to bind G actin<br />

or enter the nucleus. Both Tb4 <strong>and</strong> Tb4(AcSDKPT/4A) induced the PAI-1 gene to the same<br />

extent, while mutants Tb4(KLKKTET/7A) <strong>and</strong> Tb4(K16A) retained about 60% <strong>of</strong> the original<br />

activity. By proteomic analysis, the Ku80 subunit <strong>of</strong> ATP-dependent DNA helicase II was found<br />

to be associated with Tb4. Ku80 <strong>and</strong> Tb4 consistently co-immunoprecipitated in a complex from<br />

endothelial cells. Furthermore, downregulation <strong>of</strong> Ku80 by specific siRNA resulted in dramatic<br />

reduction in PAI-1 expression, both at the level <strong>of</strong> mRNA <strong>and</strong> protein synthesis. These data<br />

suggest that Ku80 functions as a novel receptor for Tb4, initiates signaling leading to activation <strong>of</strong><br />

PAI-1 expression by an as yet unknown mechanism, <strong>and</strong> mediates the intracellular activity <strong>of</strong> Tb4.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 87


i076i<br />

Signaling Pathway Involved in Inhibition <strong>of</strong><br />

PAI-1 Expression by CNP in Endothelial Cells<br />

Jerczynska H*, Cierniewski CS, Pawlowska Z<br />

Department <strong>of</strong> <strong>Molecular</strong> <strong>and</strong> Medical Biophysics, Medical University in Lodz, Lodz, Pol<strong>and</strong><br />

Presenting author e-mail: hanuka@zdn.am.lodz.pl<br />

We have reported previously that C-type natriuretic peptide (CNP) was an effective inhibitor <strong>of</strong><br />

PAI-1 synthesis <strong>and</strong> release from human endothelial cells. This inhibitory effect was stronger in<br />

Tumor Necrosis Factor alpha TNFa - stimulated cells.<br />

The signaling pathways required for this effect have not been elucidated. In the current study we<br />

investigated the signal transduction pathway involved in the inhibitory CNP action in human<br />

endothelial cells. We examined if CNP can modulate the known signaling pathways involved<br />

in the regulation <strong>of</strong> PAI-1 expression as well as a potential cGMP role. To characterize the signal<br />

transduction pathway, we used selective inhibitors <strong>of</strong> MAP kinase, PI3K/Akt, <strong>and</strong> 8-bromocGMP,<br />

a soluble analog <strong>of</strong> cGMP. The activation level <strong>of</strong> ERK1/2, JNK <strong>and</strong> NFaB was determined<br />

by ELISA with the use <strong>of</strong> specific antibodies. Our results imply that CNP inhibition <strong>of</strong> TNFaactivated<br />

PAI-1 gene expression takes place via regulation <strong>of</strong> signal transduction pathways<br />

involving directly PI3K/Akt <strong>and</strong> cGMP-dependent protein kinase, possibly by inhibiting NFkBdependent<br />

pathways. Thus, our results explain at least in part the mechanism involved in PAI-1<br />

expression inhibition by CNP in TNF-stimulated endothelial cells.<br />

88 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i077i<br />

Interaction <strong>of</strong> Alzheimer’s Amyloid b-peptide (Ab) 1-40 with PAI-2<br />

Pabba M* 1 , Przygodzki T 1 , Malisauskas M 1 , Ol<strong>of</strong>sson A 2 , Morozova-Roche L 1 , Wilczynska M 1 , Ny T 1<br />

1 Department <strong>of</strong> Medical Biochemistry <strong>and</strong> Biophysics, 2 Umeå Center for <strong>Molecular</strong> Pathogenesis, Umeå<br />

University, Umeå, Sweden<br />

Presenting author e-mail: pabba.mohan@medchem.umu.se<br />

Alzheimer’s disease (AD) is one <strong>of</strong> the most known dementia in elderly. The disease is<br />

characterized by the presence <strong>of</strong> senile plaques that are almost entirely composed <strong>of</strong> an amyloid<br />

b-peptide. Amyloid b-peptide which is a proteolytic fragment produced from amyloid precursor<br />

protein, is also found in intracellular compartments, <strong>and</strong> recently several physiological roles <strong>of</strong> the<br />

peptide have been proposed. Acceleration as well as inhibition <strong>of</strong> Ab-peptide protein aggregation<br />

during Alzheimer’s disease is accomplished by various factors, including serpins. Serpins are<br />

the largest <strong>and</strong> broadly distributed superfamily <strong>of</strong> protease inhibitors. <strong>Plasminogen</strong> activator<br />

inhibitor type 2 (PAI-2) belongs to ovalbumin sub-clade <strong>of</strong> the serpin superfamily. Although it<br />

has been extensively studied, its biological functions remain unknown. It has been shown that<br />

expression <strong>of</strong> PAI-2 is high in microglia cells surrounding senile plaques in the brain <strong>of</strong> patients<br />

with Alzheimer’s disease, suggesting that PAI-2 might have a role in disease. Our western<br />

blot <strong>and</strong> sucrose cushion sedimentation experiments show that Ab-peptide interacted <strong>and</strong> coaggregate<br />

with PAI-2. In addition, atomic force microscopy <strong>and</strong> thi<strong>of</strong>lavin-T analyses reveal that<br />

PAI-2 inhibit prot<strong>of</strong>ibril formation by the Ab-peptide. Taken together these data suggest that<br />

Ab-peptide 1-40 interacts with PAI-2 in vitro. Studies on interaction <strong>of</strong> Ab-peptide with PAI-2 in<br />

neuronal cells are in progress.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 89


i078i<br />

The Subcellular Itinerary <strong>of</strong> Hepatocyte Growth<br />

Factor Activator Inhibitor-1 in MDCK Cells<br />

Godiksen S 1 , Selzer-Plon J 1 , Pedersen EDK 1 , Borger Rasmussen H 1 , Bugge TH 2 , Vogel LK* 1<br />

1 Department <strong>of</strong> <strong>Cellular</strong> <strong>and</strong> <strong>Molecular</strong> Medicine, University <strong>of</strong> Copenhagen, Denmark;<br />

2 Oral <strong>and</strong> Pharyngeal Cancer Branch, NIDCR, NIH, Bethesda, Maryl<strong>and</strong>, USA<br />

Presenting author e-mail: vogel@imbg.ku.dk<br />

Hepatocyte growth factor activator inhibitor-1 (HAI-1) is a Kunitz-type transmembrane serine<br />

protease inhibitor that forms inhibitor complexes with several trypsin-like serine proteases that<br />

is required for mouse placental development <strong>and</strong> embryo survival <strong>and</strong> is a key regulator <strong>of</strong><br />

carcinogenesis. HAI-1 is expressed in polarized epithelial cells, where the plasma membrane<br />

is divided by tight juntions into an apical <strong>and</strong> a basolateral domain. Here we show that HAI-<br />

1 at steady state is mainly located on the basolateral membrane in both MDCK cells <strong>and</strong><br />

mammary gl<strong>and</strong> epithelial cells. After biosynthesis HAI-1 is exocytosed to the basolateral plasma<br />

membrane from where 20% <strong>of</strong> the HAI-1 molecules are proteolytically cleaved <strong>and</strong> released<br />

into the basolateral media. The remaining membrane associated HAI-1 is endocytosed <strong>and</strong><br />

recycles between the basolateral plama membrane <strong>and</strong> endosomes. However, a minor fraction<br />

<strong>of</strong> basolaterally located HAI-1 is trancytosed to the apical plasma membrane, from where it<br />

is probably cleaved <strong>and</strong> released into the media. HAI-1 is raft associated during most <strong>of</strong> its<br />

intracellular transport in a methyl-b-cyclodextrin dependent manner.<br />

90 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i079i<br />

Evidence for a Matriptase-Prostasin (CAP1/PRSS8) Serine Protease<br />

Zymogen-Cascade-Regulating Epithelial Differentiation<br />

List K* 1 , Netzel-Arnett S 2 , Currie B 1 , Szabo R 1 , Molinolo A 1 , Antalis TM 2 , Bugge TH 1<br />

1 Oral <strong>and</strong> Pharyngeal Cancer Branch, NIDCR, NIH, Bethesda, Maryl<strong>and</strong>, USA;<br />

2 Center for Vascular <strong>and</strong> Inflammatory Disease, University <strong>of</strong> Maryl<strong>and</strong>, Baltimore, Maryl<strong>and</strong>, USA<br />

Presenting author e-mail: klist@nidcr.nih.gov<br />

Epidermal ablation <strong>of</strong> matriptase <strong>and</strong> prostasin has identical deleterious effects on terminal<br />

differentiation, suggesting a functional interrelationship between the two membrane serine<br />

proteases. Here we present histological, biochemical, <strong>and</strong> genetic evidence for a matriptaseprostasin<br />

zymogen cascade regulating epithelial differentiation. Enzymatic gene trapping <strong>of</strong><br />

matriptase combined with prostasin immunohistochemistry revealed that matriptase was colocalized<br />

with prostasin in transitional layer cells <strong>of</strong> the epidermis, <strong>and</strong> that the developmental<br />

onset <strong>of</strong> expression <strong>of</strong> the two membrane proteases correlated with acquisition <strong>of</strong> epidermal<br />

barrier function. Purified soluble matriptase efficiently converted soluble prostasin zymogen to<br />

an active two-chain form that formed SDS-stable complexes with the serpin protease nexin-1.<br />

Whereas two forms <strong>of</strong> prostasin with molecular weights corresponding to the prostasin zymogen<br />

<strong>and</strong> active prostasin were present in wildtype epidermis, prostasin was exclusively found in<br />

the zymogen form in matriptase-deficient epidermis. Moreover, we found that matriptase <strong>and</strong><br />

prostasin displayed a near-ubiquitous co-localization in simple, stratified, <strong>and</strong> pseudo-stratified<br />

epithelia <strong>of</strong> the integumentary system, digestive tract, respiratory tract, <strong>and</strong> urogenital tract,<br />

suggesting that the proposed matriptase-prostasin zymogen cascade may have additional roles in<br />

epithelial biology besides regulating terminal epidermal differentiation. Because matriptase is an<br />

auto-activating protease, it may serve as the initiator <strong>of</strong> zymogen cascades, similar to the closely<br />

related protease, enteropeptidase.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 91


i080i<br />

The <strong>Plasminogen</strong> <strong>Activation</strong> System in Monocytic Cell<br />

Differentiation <strong>and</strong> Proliferation: Potential Target for<br />

<strong>Plasminogen</strong> <strong>Activation</strong> Inhibitor Type-2-Based Therapeutics<br />

Lee JA* 1 , Croucher DR 2 , Ranson M 1<br />

1 School <strong>of</strong> Biological Sciences, University <strong>of</strong> Wollongong, NSW Australia;<br />

2 Cancer Research Program, Garvan Institute <strong>of</strong> Medical Research, Sydney, Australia<br />

Presenting author e-mail: jal31@uow.edu.au<br />

The progression <strong>of</strong> tumours to malignancy is aided by the surrounding stroma <strong>and</strong> infiltrating<br />

leukocytes, <strong>of</strong> which tumour associated macrophages (TAMs) contribute up to 70% <strong>of</strong> the tumour<br />

mass in some cancers. TAMs potentiate malignancy by contributing to the production <strong>of</strong> the<br />

extracellular protease, plasmin. Originating from peripheral blood monocytes, TAMs have been<br />

found to over express components <strong>of</strong> the u-PA system <strong>and</strong> this behaviour identifies them as<br />

potential targets in the development <strong>of</strong> anti-u-PA targeted cancer therapies. One such potential<br />

therapy utilises the accelerated internalisation mechanisms <strong>of</strong> cell surface u-PA upon inhibition<br />

by its specific inhibitor, plasminogen activation inhibitor (PAI) type-2. This research shows for<br />

the first time that the internalisation <strong>of</strong> PAI-2 by differentiated U-937 <strong>and</strong> THP-1 monocytic cell<br />

lines is mediated in part by the low density lipoprotein receptor (LDLr) family <strong>of</strong> endocytosis<br />

receptors, in a u-PA dependent manner. Importantly, this shows that TAMs expressing high levels<br />

<strong>of</strong> u-PA may be targeted using PAI-2 therapeutics. Assessment <strong>of</strong> physiological consequences<br />

associated with exogenous PAI interactions at the cell surface revealed a reduction in proliferation<br />

for both cell lines when cultured in the presence <strong>of</strong> PAI-2 <strong>and</strong> PAI type-1, suggesting that u-PA:<br />

PAI complex formation at the cell surface disrupts mitogenic signalling cascades mediated via<br />

receptor bound u-PA. Based on these findings current research is being directed at the role <strong>of</strong><br />

exogenous PAI-2 in proliferation <strong>and</strong> differentiation <strong>of</strong> monocytes.<br />

92 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i081i<br />

Identification <strong>and</strong> Localization <strong>of</strong> Novel Serine<br />

Proteases in the Mouse Ovary<br />

Wahlberg P*, Nyl<strong>and</strong>er Å, Kui L, Ny T<br />

Department <strong>of</strong> Medical Biochemistry <strong>and</strong> Biophysics, Umeå University, Umeå, Sweden<br />

Presenting author e-mail: patrik.wahlberg@medchem.umu.se<br />

Proteolytic degradation <strong>of</strong> extracellular matrix components is believed to play an essential role<br />

for ovulation to occur. Recent studies in our laboratory have suggested that the plasminogen (PA)<br />

<strong>and</strong> matrix metalloproteinase (MMP) systems are not required in this process. In this study, we<br />

have used a microarray approach to identify new proteases that are involved in ovulation. We<br />

discovered three new serine proteases that were relatively highly expressed during ovulation:<br />

HtrA1, which was not regulated during ovulation; PRSS23, which was downregulated by<br />

gonadotropins; <strong>and</strong> PRSS35, which was upregulated by gonadotropins. We have investigated<br />

the expression patterns <strong>of</strong> these proteases during gonadotropin-induced ovulation in immature<br />

mice <strong>and</strong> in the corpus luteum (CL) <strong>of</strong> pseudopregnant mice. We found that HtrA1 was highly<br />

expressed in granulosa cells throughout follicular development <strong>and</strong> ovulation. It was also highly<br />

expressed in the forming <strong>and</strong> regressing CL. PRSS23 was highly expressed in atretic follicles <strong>and</strong><br />

it was expressed in the ovarian stroma <strong>and</strong> theca tissues after ovulation was induced. PRSS35 was<br />

expressed in the theca layers <strong>of</strong> developing follicles. It was also highly induced in granulosa cells<br />

<strong>of</strong> preovulatory follicles. PRSS35 was also expressed in the forming <strong>and</strong> regressing CL. These data<br />

suggest that HtrA1 <strong>and</strong> PRSS35 are involved in ovulation <strong>and</strong> CL formation <strong>and</strong> regression <strong>and</strong><br />

that PRSS23 may play a role in follicular atre.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 93


i082i<br />

Alpha-enolase/<strong>Plasminogen</strong> Binding Is Required<br />

during Myogenesis in vitro <strong>and</strong> in vivo<br />

Diaz-Ramos A*, Llorens A, Luque T, López-Alemany R<br />

Institute <strong>of</strong> Biomedical Investigation <strong>of</strong> Bellvitge (IDIBELL), <strong>Molecular</strong> Oncology Center (COM),<br />

L’Hospitalet de Llobregat, Barcelona, Spain<br />

Presenting author e-mail: madiaz@idibell.org<br />

During myogenesis, an important proteolyitc activity <strong>and</strong> extensive extracellular matrix (ECM)<br />

remodelation takes place. Plasmin, generated by activation <strong>of</strong> plasminogen, is an extracellular<br />

protease specialized in the degradation <strong>of</strong> the ECM components. Alpha-enolase constitutes a<br />

receptor for plasminogen in several cell lines, serving to focalize proteolytic activity on the cell<br />

surface.<br />

Previous studies show a role for components <strong>of</strong> the plasminogen activation system during<br />

myogenesis in vitro <strong>and</strong> in vivo. Preliminary results <strong>of</strong> our laboratory have shown that alphaenolase<br />

expression was up-regulated in a myogenic cell line upon differentiation, but the role <strong>of</strong><br />

alpha-enolase as a plasminogen receptor in myogenesis deserves further analysis.<br />

In the presence <strong>of</strong> inhibitors <strong>of</strong> plasminogen/alpha-enolase binding, myogenic differentiation,<br />

fusion <strong>and</strong> migration were abrograted, using primary cultures <strong>of</strong> myogenic cells or Muscle<br />

Precursor Cells (MPCs).<br />

The effect <strong>of</strong> inhibitors <strong>of</strong> plasminogen/alpha-enolase binding were evaluated in a regeneration<br />

model in mice after a muscle. Regeneration parameters were blocked by inhibitors <strong>of</strong><br />

plasminogen/alpha-enolase. The mdx mice (the animal model for Duchenne Muscular Dystrophy,<br />

DMD) presented a more severe dystrophinopathy with the same treatment. Inflammatory cells<br />

infiltration <strong>and</strong> fibrin deposition in injured tissues are currently being evaluated.<br />

Since inhibitors <strong>of</strong> plasminogen/alpha-enolase binding have an inhibitory effect on MPCs<br />

differentiation <strong>and</strong> muscle regeneration in vivo, our results demonstrate that plasmin activity<br />

is necessary for myogenesis to take place correctly, in an alpha-enolase dependent way.<br />

<strong>Plasminogen</strong>/alpha-enolase binding therefore could be an important target in the development <strong>of</strong><br />

treatments for DMD.<br />

94 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i083i<br />

The Serpinb8 Is Alternatively Spliced to the Known<br />

Long Form <strong>and</strong> a Novel Short Form<br />

Olausson B* 1 , Przygodzka P 1 , Dahl L 2 , Carlsson L 2 , Wilczynska M 1<br />

1 2 Department <strong>of</strong> Medical Biochemistry <strong>and</strong> Biophysics, Umeå Center for <strong>Molecular</strong> Medicine, Umeå<br />

University, Umeå, Sweden<br />

Presenting author e-mail: bjorn.olausson@medchem.umu.se<br />

Serpinb8 belongs to the ovoalbumin-serpin subfamily, <strong>and</strong> is known as an intracellular protease<br />

inhibitor with a unique distribution pattern in human tissues. Here we show that serpinb8 can<br />

be alternatively spliced to the already known long form <strong>and</strong> to a novel short form. Compared to<br />

the long form, the short form <strong>of</strong> serpinb8 is missing exon 7, <strong>and</strong> instead it has a short nucleotide<br />

extension. Exon 7 in the long serpinb8 encodes part <strong>of</strong> beta-sheet A <strong>and</strong> the reactive center loop.<br />

Therefore, the short form <strong>of</strong> serpinb8 can be predicted not to function as protease inhibitor. By<br />

using real-time PCR, we found that in myelomonocytic leukemia cell lines, as well as in mouse<br />

hematopoietic stem cells, the long form <strong>of</strong> serpinb8 is expressed at very low levels, but the<br />

expression increases during monocytic, granulocytic, <strong>and</strong> megakaryocytic differentiations to<br />

reach maximum in finally differentiated blood cells. In contrast, the short form <strong>of</strong> serpinb8 is<br />

temporarily expressed during monocytic differentiation <strong>of</strong> HL-60 cells <strong>and</strong> is not detectable in the<br />

differentiated monocytes. Our preliminary real-time PCR data on the RNA from 18 human tissues<br />

suggest that the short form <strong>of</strong> serpinb8 is expressed only in few tissues as pancreas, oesophagus,<br />

<strong>and</strong> kidney. Studies on the cellular localization <strong>of</strong> the spliced forms <strong>of</strong> serpinb8 are in progress.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 95


i084i<br />

Characterization <strong>of</strong> a Combined PAI-1 <strong>and</strong> TIMP-1<br />

Gene-deficient Mouse Model<br />

Harslund J* 1 , Nielsen OL 2 , Brünner N 1 , Offenberg H 1<br />

1 2 Section <strong>of</strong> Biomedicine, Section <strong>of</strong> Pathology, Department <strong>of</strong> Veterinary Pathobiology, Faculty <strong>of</strong> Life<br />

Sciences, University <strong>of</strong> Copenhagen, Frederiksberg, Denmark<br />

Presenting author e-mail: jhar@life.ku.dk<br />

The endogenous proteinase inhibitors PAI-1 <strong>and</strong> TIMP-1 are two distinct proteins with separate<br />

molecular pathways. However, a much closer relationship between PAI-1 <strong>and</strong> TIMP-1 has been<br />

proposed indicating some degree <strong>of</strong> functional overlap due to their involvement in ECM turnover,<br />

tissue remodelling <strong>and</strong> cellular migration <strong>and</strong> signalling. To study the physiological implications<br />

<strong>of</strong> PAI-1 <strong>and</strong> TIMP-1 we have generated a combined PAI-1 <strong>and</strong> TIMP-1 gene deficient mouse<br />

model. We present the results on generating this specific mouse model with particular emphasis<br />

on phenotypical characteristics, haematological parameters, a histological evaluation <strong>and</strong> gene<br />

expression studies <strong>of</strong> PAI-1 <strong>and</strong> TIMP-1 in various organs. We observed a significant deviation in<br />

segregation <strong>of</strong> <strong>of</strong>fspring only in male mice (p


i085i<br />

Dual Role <strong>of</strong> the uPA/uPAR System in Apoptosis <strong>of</strong><br />

Mesangial Cells <strong>and</strong> Diabetic Nephropathy<br />

Tkachuk N, Tkachuk S*, Kiyan J, Shushakova N, Haller H, Dumler I<br />

Nephrology Department, Hannover Medical School, Hannover, Germany<br />

Presenting author e-mail: Tkatchouk.Sergei@mh-hannover.de<br />

In diabetic nephropathy apoptosis has been considered as a major mechanism for regulation<br />

<strong>of</strong> cell amount in the places <strong>of</strong> mesangial hypercellularity. Given these functions for apoptosis<br />

together with the observation that uPAR <strong>and</strong> uPA may be upregulated during renal injury, the<br />

present study was designed to investigate the role <strong>of</strong> the uPA/uPAR system in apoptosis <strong>of</strong><br />

mesangial cells (MC) under high glucose <strong>and</strong> growth factors withdrawal conditions. Our data<br />

show that uPA can elicit both pro-apoptotic <strong>and</strong> anti-apoptotic effects in human MC depending<br />

on the apoptosis-inducing stimulus. Thus, uPA abrogated MC apoptosis induced by growth<br />

factors withdrawal conditions, whereas apoptosis initiated in MC by high glucose was enhanced<br />

in presence <strong>of</strong> uPA. Amino-terminal fragment (ATF) <strong>of</strong> uPA completely retained effects <strong>of</strong> uPA.<br />

uPAR was shown to be involved in both pro-apoptotic <strong>and</strong> anti-apoptotic effects <strong>of</strong> uPA. <strong>Cellular</strong><br />

redistribution <strong>of</strong> uPAR in response to high glucose was observed. Members <strong>of</strong> MAPK <strong>and</strong> PI3kinase<br />

pathways <strong>and</strong> Bad protein were found to be downstream molecules mediating the effects<br />

<strong>of</strong> uPA. The opposite effects <strong>of</strong> uPA on MC apoptosis are directed via different uPAR-interacting<br />

transmembrane partners. Thus, whereas the anti-apoptotic effect <strong>of</strong> uPA was mediated by<br />

integrins, its pro-apoptotic effect required uPAR interaction with cation-independent mannose-<br />

6-phosphate receptor (M6PR). In vivo, uPAR <strong>and</strong> M6PR were upregulated in the kidney <strong>of</strong> mice<br />

in streptozotocin (STZ)-induced diabetes that confirmed an important role for these receptors in<br />

mediating functional response <strong>of</strong> mesangial cells to high glucose.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 97


i086i<br />

Phenotypic Consequences <strong>of</strong> <strong>Plasminogen</strong> Activator<br />

Inhibitor-1 Gene Ablation on STAT1 <strong>Activation</strong> <strong>and</strong> Cell<br />

Cycle Progression in Proliferating Endothelial Cells<br />

Balsara RD* 1,2,3 , Morin SJ 1 , Meyer CA 1 , Castellino FJ 1,2,3 , Ploplis VA 1,2,3<br />

1 2 3 W. M. Keck Center for Transgene Research, Department <strong>of</strong> Chemistry <strong>and</strong> Biochemistry, Notre Dame<br />

Cancer Institute, University <strong>of</strong> Notre Dame, Indiana, USA<br />

Presenting author e-mail: rbalsara@nd.edu<br />

The enhanced proliferation observed in PAI-1-deficient endothelial cells (EC) is associated with<br />

Akt hyperactivation resulting in loss <strong>of</strong> growth control <strong>and</strong> resistance to apoptosis. Additionally,<br />

involvement <strong>of</strong> other signaling pathways that regulate cell proliferation was investigated.<br />

Evaluation <strong>of</strong> the STAT protein pr<strong>of</strong>ile revealed that PAI-1–/– EC exhibited lower levels <strong>of</strong><br />

STAT1, which is a negative regulator <strong>of</strong> cellular proliferation. The pro-proliferative STAT3 <strong>and</strong><br />

STAT5 protein levels were similar in WT <strong>and</strong> PAI-1–/– cells. Furthermore, the activation status <strong>of</strong><br />

STAT1 between WT <strong>and</strong> PAI-1–/– EC differs. The extent <strong>of</strong> phosphorylation <strong>of</strong> Ser727 within the<br />

STAT1 transcriptional activation domain is lower in PAI-1–/– EC. Immun<strong>of</strong>luorescent analyses<br />

<strong>of</strong> proliferating WT <strong>and</strong> PAI-1–/– EC demonstrate that the cellular distribution <strong>of</strong> STAT1(P-<br />

Tyr701), which is crucial for STAT1 dimerization <strong>and</strong> nuclear translocation, is different between<br />

the two cell types. In WT cells, STAT1(P-Tyr701) appears to be diffuse throughout the whole<br />

cell, including the nucleus. In contrast, in PAI-1–/– cells, STAT1(P-Tyr701) seems to be localized<br />

in the cytoplasm <strong>and</strong> plasma membrane. Additionally, it was observed that in WT cells STAT1<br />

preferentially binds to Jak1, whereas in the PAI-1–/– EC, STAT1 is preferentially bound to Jak2.<br />

This differential binding may be responsible for the different activation states <strong>of</strong> STAT1. Cell cycle<br />

progression analysis demonstrated that a higher percentage <strong>of</strong> PAI-1–/– EC are in the S-phase<br />

compared to WT cells, which is likely a consequence <strong>of</strong> diminished STAT1 levels observed in the<br />

PAI-1-deficient cells.<br />

98 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i087i<br />

Urokinase <strong>and</strong> its Receptor as Novel C-Myc Target<br />

Genes Affecting Cell Migration <strong>and</strong> Apoptosis<br />

Alfano D*, Iaccarino I, Stoppelli MP<br />

Institute <strong>of</strong> Genetics <strong>and</strong> Biophysics ‘’Adriano Buzzati-Traverso’’, National Research Council, Naples,<br />

Italy<br />

Presenting author e-mail: alfano@igb.cnr.it<br />

The binding <strong>of</strong> the serine protease urokinase (uPA) to its receptor (uPAR) plays a central role<br />

in cell adhesion, migration <strong>and</strong> invasion. According to our recently published data, uPAR<br />

overexpression <strong>and</strong> signalling prevents DNA damage-induced apoptosis as well as anoikis <strong>of</strong><br />

retinal pigment epithelial cells (RPE). Among the genes regulating cell proliferation <strong>and</strong> survival<br />

is the oncogene c-Myc. In an attempt to characterise the effect <strong>of</strong> c-Myc activation on uPA <strong>and</strong><br />

uPAR expression as well as on the cell survival/apoptosis balance, we employed an RPE cell line<br />

carrying an inducible c-MycER. As previously reported in other cell systems, induction <strong>of</strong> c-Myc<br />

in RPE cells leads to an increased cell sensitivity to different pro-apototic stimuli, like uv-light<br />

<strong>and</strong> serum starvation. To investigate the genes regulated in these conditions, the gene expression<br />

pr<strong>of</strong>ile following c-Myc activation for 4-16 hrs was determined. Surprisingly, a marked downregulation<br />

<strong>of</strong> the mRNAs coding for uPA <strong>and</strong> uPAR was observed. Furthermore, addition <strong>of</strong><br />

recombinant uPA to the culture medium improuves cell survival, suggesting that<br />

c-Myc-induced apoptosis is, at least in part, due to uPA down-regulation. As expected, following<br />

c-Myc activation, we observed a remarkable decrease in cell migration ability. In conclusion, the<br />

uPA/uPAR system is subjected to a c-Myc-dependent downregulation under conditions in which<br />

a concomitant increase in the sensitivity to apoptosis <strong>and</strong> a decrease in cell motility is observed.<br />

We propose that during oncogenic transformation the uPA/uPAR system might restrain c-Myc<br />

proliferative <strong>and</strong> invasive potential.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 99


i088i<br />

Effect <strong>of</strong> <strong>Plasminogen</strong> on Cell Migration Using an in vitro Wound Model<br />

Sulniute R*, Li J, Ny T<br />

Department <strong>of</strong> Medical Biochemistry <strong>and</strong> Biophysics, Umeå University, Umeå, Sweden<br />

Presenting author e-mail: rima.sulniute@medchem.umu.se<br />

In vivo studies on incisional wounds have shown that plasminogen (plg) plays an important role<br />

in wound healing. However, due to the complexity <strong>of</strong> the in vivo system, the exact functional roles<br />

that plg plays in wound healing at the molecular level remain obscure.<br />

We therefore established a simplified in vitro model in order to directly investigate the molecular<br />

mechanism by which plg affects keratinocyte migration in wound healing. In this model, a<br />

scratch wound is made on a monolayer <strong>of</strong> adherent DOK (early neoplastic/dysplastic human oral<br />

keratinocytes) cell line <strong>and</strong> the wound closure process is monitored thereafter. Our data reveal<br />

that at 24 hours after scratch wounding, DOK cells had migrated about 65% <strong>of</strong> the original wound<br />

area in the presence <strong>of</strong> 2 µM plg (mimicking the plasma level <strong>of</strong> plg), whereas these cells only<br />

migrated about 45% when incubated in the plg-depleted medium. Furthermore, wound closure<br />

was accelerated up to 80% <strong>of</strong> the original wound when the level <strong>of</strong> plg was doubled<br />

(4 µM). Interestingly, when PMSF, a serine protease inhibitor, was added in the presence <strong>of</strong> plg,<br />

the cells had the same migration rate as those in the plg-depleted medium. Altogether, our data<br />

suggest that plg, probably through its proteolytic activity, is essential in keratinocyte migration in<br />

an in vitro wound healing model. These findings are in consistence with our in vivo studies in plg<br />

deficient mice.<br />

100 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i089i<br />

uPA, but not its Receptor uPAR, Is Necessary for Experimentallyinduced<br />

<strong>and</strong> Pathological Muscle Regeneration<br />

Vidal B*, Serrano AL, Jardí M, Suelves M, Muñoz-Cánoves P<br />

Center for Genomic Regulation (CRG), Differentiation <strong>and</strong> Cancer Program, Barcelona, Spain<br />

Presenting author e-mail: berta.vidal@crg.es<br />

Extracellular proteolysis takes place during skeletal muscle regeneration. We have previously<br />

shown that uPA is required for muscle regeneration in vivo, since regeneration was impaired<br />

in uPA–/– mice after muscle injury. Moreover, the double mutant uPA–/–mdx mice showed<br />

exacerbated dystrophinopathy compared to uPA+/+mdx mice, the animal model for Duchenne<br />

muscular dystrophy. uPA was critical for a correct inflammatory response, as demonstrated by<br />

the reduced number <strong>of</strong> infiltrated macrophages <strong>and</strong> T-cells in uPA–/–mdx mice, as compared<br />

to uPA+/+mdx mice. Consistent herewith, loss <strong>of</strong> uPA also reduced the number <strong>of</strong> infiltrated<br />

inflammatory cells in experimentally-injured muscle. In contrast, the muscular compartment<br />

did not seem to be affected by loss <strong>of</strong> uPA, since isolated uPA–/– satellite cells showed normal<br />

myogenesis in vitro. Since several cellular functions <strong>of</strong> uPA do not require its proteolytical activity,<br />

but rather its ability to bind uPAR, <strong>and</strong> because uPAR expression was induced in regenerating<br />

muscle <strong>of</strong> WT <strong>and</strong> mdx mice, we hypothesized that the role <strong>of</strong> uPA in muscle regeneration might<br />

be dependent, at least in part, on its binding to uPAR. Thus, we analyzed the consequences <strong>of</strong><br />

uPAR deficiency (uPAR–/– mice) in experimentally-induced muscle regeneration <strong>and</strong> mdx<br />

dystrophinopathy (uPAR–/–mdx mice). Experimentally-induced muscle regeneration was<br />

indistinguishable between WT <strong>and</strong> uPAR–/– mice after histological analyses at 2, 10 <strong>and</strong> 25<br />

days post-injury. Muscular dystrophy was also similar in uPAR+/+mdx <strong>and</strong> uPAR–/–mdx<br />

mice, indicating that uPAR is dispensable for muscle tissue remodeling during regeneration.<br />

These results suggest that uPA regulates key processes during muscle regeneration in a uPARindependent<br />

manner.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 101


i090i<br />

Urokinase-type <strong>Plasminogen</strong> Activator Deficiency Strongly Attenuates<br />

Ischemia Reperfusion Injury <strong>and</strong> Acute Kidney Allograft Rejection<br />

Gueler F 1 , Rong S 1 , Mengel M 2 , Park J-K 1 , Kirsch T 1 , Haller H 1 , Dumler I 1 , Shushakova N* 1<br />

1 Department <strong>of</strong> Nephrology, 2 Department <strong>of</strong> Pathology, Medical School Hanover, Hanover, Germany<br />

Presenting author e-mail: nshushakova@phenos.com<br />

Central mechanisms leading to ischemia induced allograft rejection are apoptosis <strong>and</strong><br />

inflammation, processes highly regulated by the urokinase-type plasminogen activator (uPA)<br />

<strong>and</strong> its specific receptor (uPAR). Recently, up-regulation <strong>of</strong> uPA <strong>and</strong> uPAR has been shown to<br />

correlate with allograft rejection in human biopsies. However, the causal connection <strong>of</strong> uPA/<br />

uPAR in mediating <strong>of</strong> transplant rejection <strong>and</strong> underlying molecular mechanisms remain poorly<br />

understood. In this study, we evaluated the role <strong>of</strong> uPA/uPAR in a mice model for IR injury<br />

<strong>and</strong> for acute kidney allograft rejection. uPAR but not uPA deficiency strongly protected kidney<br />

from IR injury. For the transplant model, allografts from H2b uPA–/–, uPAR–/– <strong>and</strong> wildtype<br />

(WT) mice were transplanted into MHC-incompatible H2d WT recipients. Again uPAR but not<br />

uPA deficiency <strong>of</strong> the allograft caused superior recipient survival <strong>and</strong> strongly attenuated loss <strong>of</strong><br />

renal function. We could demonstrate that uPAR-deficient allografts showed reduced generation<br />

<strong>of</strong> reactive oxygen species (ROS) <strong>and</strong> apoptosis. Moreover, transplant-induced up-regulation<br />

<strong>of</strong> adhesion molecule ICAM-1 was completely abrogated in uPAR-deficient allografts due to<br />

inadequate C5a <strong>and</strong> TNFa signalling, leading to reduced leukocyte infiltration. Our results<br />

demonstrate that the local renal uPAR plays an important role in the apoptotic <strong>and</strong> inflammatory<br />

responses mediating IR-injury <strong>and</strong> transplant rejection.<br />

102 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i091i<br />

a1-antitrypsin Polymerization Studies using Gas-phase<br />

Electrophoretic Mobility <strong>Molecular</strong> Analysis (GEMMA)<br />

Przygodzki T* 1 , Mallya M 2 , Phillips RL 2 , Belorgey D 2 , Hägglöf P 2 , Lomas DA 2 , Ny T 1<br />

1 Department <strong>of</strong> Medical Biochemistry <strong>and</strong> Biophysics, Umeå University, Umeå, Sweden;<br />

2 Department <strong>of</strong> Medicine, University <strong>of</strong> Cambridge, Cambridge Institute for Medical Research,<br />

Cambridge, United Kingdom<br />

Presenting author e-mail: tomasz.przygodzki@medchem.umu.se<br />

Several proteins from the serpin superfamily are known to have capability <strong>of</strong> forming polymers.<br />

One <strong>of</strong> them is the elastase inhibitor-a1-antitrypsin (AT). Beside wild type AT which does not<br />

polymerize in physiological conditions, a highly polymerizing, natural mutant, Z AT, has been<br />

described.<br />

The kinetics <strong>of</strong> serpin polymerization is still a matter <strong>of</strong> debate. We followed this process by a<br />

novel approach: gas-phase electrophoretic mobility molecular analysis (GEMMA). This technique<br />

allows the analysis <strong>of</strong> the distribution <strong>of</strong> macromolecules in solution with respect to their size. A<br />

relatively short time required for sample analyses, a broad range <strong>of</strong> macromolecule size that can<br />

be separated as well as the fact that non-covalent complexes remain intact upon measurement<br />

make this technique very useful for the study <strong>of</strong> protein assembly.<br />

We followed the polymerization process <strong>of</strong> both wild type <strong>and</strong> Z AT over time at three<br />

temperatures. We observed no oligomers formation <strong>of</strong> wild type AT at temperatures <strong>of</strong> 36°C <strong>and</strong><br />

45°C whereas Z AT was polymerizing in all conditions. Velocity <strong>of</strong> wt AT oligomers formation<br />

at 55°C was approximately 10 times lower than that for the Z mutant. We have observed<br />

dimeric, trimeric, tetrameric <strong>and</strong> pentameric species in the case <strong>of</strong> both proteins indicating that<br />

polymerization proceeds, by sequential addition <strong>of</strong> monomers, however the fast formation <strong>of</strong><br />

dimers suggests that the dimer can also be a building block in formation <strong>of</strong> the polymers.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 103


i092i<br />

The <strong>Plasminogen</strong> Interaction <strong>of</strong> Antigen 85B Protein<br />

from Mycobacterium Tuberculosis: Role <strong>of</strong> Lys89<br />

Xolalpa W* 1 , Vallecillo AJ 1 , Rosales L 2 , Ruiz BH 2 , Espitia C 1<br />

1 Departamento de Inmunología y, 2 Departamento de Biología <strong>Molecular</strong>, Instituto de Investigaciones<br />

Biomédicas, Universidad Nacional Autónoma de México, México, México<br />

Presenting author e-mail: wendyxv@hotmail.com<br />

Interactions <strong>of</strong> microorganims with the <strong>Plasminogen</strong> (Plg) system molecules have been<br />

recognized in numerous bacteria species. Our group has recently identified several Plg binding<br />

proteins in Mycobacterium tuberculosis extracts combining 2D-PAGE <strong>and</strong> Lig<strong>and</strong> Blotting tools.<br />

Interestingly, proteins belonging to the antigen 85 (Ag85) complex, which have a role on cell wall<br />

synthesis <strong>and</strong> bind the extracellular matrix protein fibronectin, were identified as Plg binding<br />

proteins. This binding was inhibited by the lysine analogue e-aminocaproic acid, involving<br />

the participation <strong>of</strong> lysine binding sites (LBSs) <strong>of</strong> Plg kringles. In order to characterize the Plg-<br />

Ag85 interaction, we took advantage <strong>of</strong> molecular modeling tools. Using the theoretical model<br />

<strong>of</strong> the lysine-LBS interaction <strong>of</strong> the homodimeric complex <strong>of</strong> kringle 2 from tPA as reference <strong>of</strong><br />

interactions between a kringle domain <strong>and</strong> a non C-terminal lysine, the Lys89 from Ag85A <strong>and</strong><br />

Ag85B was recognized as a better c<strong>and</strong>idate to mediate Plg interaction. To assess the role <strong>of</strong> this<br />

aa, site-specific mutation was introduced in the M. tuberculosis Ag85B recombinant protein<br />

replacing the lysine by arginine. Recombinant Ag85B <strong>and</strong> Ag85BArg89 proteins were evaluated<br />

in their capability to bind to Plg by a solid-phase binding assay on ELISA plates. The results<br />

demonstrate that Plg binding on Ag85BArg89 protein is reduced until 90 percent in comparison<br />

with the wild type Ag85B protein. Binding to fibronectin was also evaluated in both proteins but<br />

no significant difference was found. These results confirm the specific role <strong>of</strong> Lys89 on Ag85B-Plg<br />

interaction giving to Ag85B a new role as putative Plg receptor.<br />

104 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i093i<br />

<strong>Plasminogen</strong> as a Factor in Innate Immunity<br />

Ahlskog N*, Guo Y, Ny T<br />

Department <strong>of</strong> Medical Chemistry <strong>and</strong> Biophysics, Umeå University, Umeå, Sweden<br />

Presenting author e-mail: nina.ahlskog@medchem.umu.se<br />

Recent results in our group strongly indicate a link between plasmin/plasminogen <strong>and</strong> the innate<br />

immune system, the body’s initial response to infection. <strong>Plasminogen</strong>-deficient mice were shown<br />

to have a higher mortality during S. aureus-induced infection (i. v. injection <strong>of</strong> low dose viable<br />

bacteria), but had a better survival rate during the initial stages <strong>of</strong> sepsis (high dose bacteria) as<br />

compared to their wild-type littermates. This phenotypic switch was shown to be specific to the<br />

active plasmin, as well as dependent on plasmin levels.<br />

Using a human mast cell line (HMC-1.2) <strong>and</strong> a mouse intra-peritoneal macrophage cell line (J774),<br />

we have found that the phenotypic switch seen in vivo, is not due to the phagocytizing capability<br />

<strong>of</strong> the macrophages or the anti-microbial activity <strong>of</strong> the mast cells in response to infection <strong>and</strong><br />

sepsis. The cells were exposed to viable bacteria in doses mimicking infection <strong>and</strong> sepsis for 20<br />

hours, after which the bacterial number, as well as the cell number, was ascertained. The antimicrobial<br />

response <strong>of</strong> both the cell types was comparable, regardless <strong>of</strong> the presence or absence<br />

<strong>of</strong> plasminogen in the medium. The migrating capability <strong>of</strong> the cells was tested using Boyden<br />

Micro Chemotaxis Chamber, where the percentage <strong>of</strong> cells migrating through the membrane <strong>of</strong><br />

the chamber was found to be comparable for both cell types, with or without plasminogen in the<br />

medium. The cellular response, e.g. release <strong>of</strong> cytokines <strong>and</strong> proteases, <strong>of</strong> the cells upon exposure<br />

to bacteria will be studied further as well as other leukocytic cell types.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 105


i094i<br />

The Inflammatory Cytokine Oncostatin M Induces <strong>Plasminogen</strong><br />

Activator Inhibitor-1 in Human Vascular Smooth Muscle Cells in<br />

vitro via PI3 kinase <strong>and</strong> MAP-kinase Dependent Pathways<br />

Demyanets S 1 , Kaun C 1 , Rychli K 1 , Rega G 1 , Pfaffenberger S 1 , Maurer G 1 , Huber K 2 , Wojta J* 1<br />

1Department <strong>of</strong> Internal Medicine II, Medical University <strong>of</strong> Vienna, Ludwig Boltzmann Foundation for<br />

Cardiovascular Research, Vienna, Austria;<br />

2 3rd Medical Department for Cardiology <strong>and</strong> Emergency Medicine, Wilhelminenspital, Vienna, Austria<br />

Presenting author e-mail: johann.wojta@meduniwien.ac.at<br />

<strong>Plasminogen</strong> activator inhibitor-1 (PAI-1) plays a pivotal role in the regulation <strong>of</strong> the fibrinolytic<br />

system <strong>and</strong> in the modulation <strong>of</strong> extracellular proteolysis. Increased expression <strong>of</strong> PAI-1 was<br />

found in atherosclerotic lesions <strong>and</strong> high PAI-1 plasma levels were shown to be associated<br />

with coronary heart disease. Smooth muscle cells seem to be a major source <strong>of</strong> PAI-1 within<br />

the vascular wall <strong>and</strong> PAI-1 was implicated in smooth muscle cell migration, proliferation <strong>and</strong><br />

apoptosis. We treated human coronary artery smooth muscle cells (HCASMC) <strong>and</strong> human aortic<br />

SMC (HASMC) with the glycoprotein 130 (gp130) lig<strong>and</strong>s cardiotrophin-1 (CT-1), interleukin-<br />

6 (IL-6), leukemia inhibitory factor (LIF) or oncostatin M (OSM). OSM, but not CT-1, IL-6 or<br />

LIF increased PAI-1 production significantly in both HCASMC <strong>and</strong> HASMC up to 20-fold as<br />

determined by a specific ELISA. OSM upregulated also mRNA specific for PAI-1 up to 4.5-fold<br />

after 24 hours in these cells <strong>and</strong> HCASMC <strong>and</strong> HASMC were shown to express gp130, OSM<br />

receptor, IL-6 receptor <strong>and</strong> LIF receptor using RealTime-polymerase chain reaction (RealTime-<br />

PCR). PD98059, a MEK inhibitor <strong>and</strong> LY294002, a PI3K inhibitor, but not AG-490, a JAK/STAT<br />

inhibitor, abolished the OSM-dependent PAI-1 induction almost completely. We hypothesize that<br />

if the effect <strong>of</strong> OSM on PAI-1 expression in smooth muscle cells is operative in vivo it could—via<br />

modulation <strong>of</strong> fibrinolysis <strong>and</strong> extracellular proteolysis—be involved in the development <strong>of</strong><br />

vascular pathologies such as plaque progression, destabilization <strong>and</strong> subsequent thrombus<br />

formation, <strong>and</strong> restenosis <strong>and</strong> neointima formation.<br />

106 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i095i<br />

Crohn’s Disease but not Chronic Ulcerative Colitis<br />

Induces the Expression <strong>of</strong> PAI-1 in Enteric Neurons<br />

Laerum OD* 1,2 , Illemann M 1 , Skarstein A 3 , Helgel<strong>and</strong> L 2 , Øvrebø K 3 , Danø K 1 , Nielsen BS 1<br />

1 The Finsen Laboratory, Rigshospitalet, Copenhagen, Denmark;<br />

2The Gade Institute, Section <strong>of</strong> Pathology, University <strong>of</strong> Bergen <strong>and</strong> Department <strong>of</strong> Pathology, Haukel<strong>and</strong><br />

University Hospital, Bergen, Norway;<br />

3Department <strong>of</strong> Surgical Sciences, University <strong>of</strong> Bergen, <strong>and</strong> Department <strong>of</strong> Surgery, Haukel<strong>and</strong><br />

University Hospital, Bergen, Norway<br />

Presenting author e-mail: ole.laerum@gades.uib.no<br />

Chronic inflammation <strong>of</strong> the intestinal wall is the common characteristic <strong>of</strong> Crohn’s disease<br />

<strong>and</strong> ulcerative colitis; disorders, which in some cases can be difficult to distinguish. The<br />

inflammation also affects the local neuronal plexuses <strong>of</strong> the enteric nervous system. It is known<br />

that plasminogen activator inhibitor-1 (PAI-1) <strong>and</strong> urokinase receptor (uPAR) are upregulated in<br />

neurons after experimental peripheral nerve injury <strong>and</strong> have been linked to nerve regeneration.<br />

The expression <strong>of</strong> PAI-1 <strong>and</strong> uPAR in neuronal cells in lesions <strong>of</strong> the gastrointestinal tract was<br />

analyzed by immunohistochemical techniques. PAI-1 was found in a subset <strong>of</strong> neurons primarily<br />

located in the submucosal plexus <strong>of</strong> the small <strong>and</strong> large intestine in 24 <strong>of</strong> 28 cases (86%) with<br />

Crohn’s disease, but in none <strong>of</strong> 17 cases with chronic ulcerative colitis <strong>and</strong> not in normal colon<br />

(n=5), acute appendicitis (n=9) or colon adenocarcinomas (n=17). The PAI-1 was seen in the<br />

perikarya <strong>of</strong> the neurons <strong>and</strong> a few proximal axons, whereas nerves were negative. uPAR was<br />

seen in nerves in all types <strong>of</strong> lesion varying from 21-88% <strong>of</strong> the cases, most frequent in colon<br />

adenocarcinomas. No uPAR-positive nerves were detected in normal colon. PAI-1-positive<br />

neurons in inflammatory bowel disease are linked to chronic inflammation in Crohn’s disease,<br />

implying PAI-1 as a potential parameter for the differential diagnosis between Crohn’s disease<br />

<strong>and</strong> ulcerative colitis. The findings also suggest that PAI-1 in neurons is related to pain <strong>and</strong> that<br />

both PAI-1 <strong>and</strong> uPAR are involved in neuronal repair in the inflamed tissue.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 107


i096i<br />

The Effect <strong>of</strong> Matrix Metalloprotease 3 Deficiency<br />

in Spontaneous Metastasis<br />

Juncker-Jensen A*, Rømer J, Almholt K<br />

Finsenlaboratoriet, Rigshospitalet, København, Denmark<br />

Presenting author e-mail: ajjensen@finsenlab.dk<br />

Cancer metastasis is dependent on proteolytic degradation <strong>of</strong> extracellular matrix barriers,<br />

making extracellular proteolysis an attractive target for anti-cancer therapy. The matrix<br />

metalloprotease (MMP) family consists <strong>of</strong> 23 (human) MMPs that are involved in physiological<br />

<strong>and</strong> pathological degradation <strong>of</strong> extracellular matrix. Several MMPs are expressed in human<br />

breast cancer, <strong>and</strong> we have demonstrated that a number <strong>of</strong> MMPs are expressed in tumors in<br />

the MMTV-PyMT transgenic breast cancer model in mice. MMP-3 (stromelysin 1) is expressed<br />

in many parts <strong>of</strong> the stroma <strong>of</strong> the tumors, resembling the expression <strong>of</strong> MMP-3 in fibroblasts <strong>of</strong><br />

human breast cancers.<br />

We have now studied the effect <strong>of</strong> MMP-3 deficiency in MMTV-PyMT mice, in a cohort <strong>of</strong> 63<br />

mice consisting <strong>of</strong> 32 MMTV-PyMT,MMP3 wild-type mice <strong>and</strong> 31 MMTV-PyMT,MMP3 knockout<br />

mice. Primary tumor growth was monitored by weekly measurements <strong>of</strong> the tumor volume<br />

<strong>and</strong> lung metastasis was quantitated using unbiased, stereological methods. All mice were<br />

sacrificed at age 87 +/– 3 days. There was no difference in tumor onset or final tumor burden<br />

when comparing wild-type <strong>and</strong> MMP-3 deficient mice. The median lung metastasis volume was<br />

increased from 0,04 mm3 in wild-type mice to 0,13 mm3 in MMP-3 deficient mice. Due to large<br />

variations in the lung metastasis volumes, this increase is not significant (Mann-Whitney U test,<br />

p=0,40).<br />

We are currently analyzing lymph node metastasis in the same mice, <strong>and</strong> we are analyzing the<br />

tumor activity <strong>of</strong> potential MMP-3 targets including the metastasis-associated MMP-2 <strong>and</strong> -9.<br />

108 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i097i<br />

Tumor-Cell Expression <strong>of</strong> C4.4A, a Structural Homologue<br />

<strong>of</strong> the Urokinase Receptor, Correlates with Poor<br />

Prognosis in Non-Small Cell Lung Cancer<br />

Skov BG 1 , Hansen LV 2 , Ploug M 2 , Pappot H* 2<br />

1 2 KAS Herlev, Department <strong>of</strong> Pathology, div. Gent<strong>of</strong>te, The Finsen Laboratory, Copenhagen University<br />

Hospital, Copenhagen, Denmark<br />

Presenting author e-mail: pappot@rh.regionh.dk<br />

C4.4A is a metastasis associated protein. This glycolipid-anchored protein is a structural<br />

homologue <strong>of</strong> the urokinase receptor, uPAR, which is a well-established prognostic marker in<br />

various human cancers. Although, several pilot studies demonstrate C4.4A expression in various<br />

human tumors, little is known about its prognostic significance. We therefore aim to explore<br />

the possible association between C4.4A expression <strong>and</strong> clinicopathological features <strong>and</strong> disease<br />

prognosis in patients with non-small cell lung cancer (NSCLC).<br />

Tissue sections from 109 NSCLC patients were subjected to immunohistochemical staining using<br />

a polyclonal antibody that specifically recognizes human C4.4A. Staining frequency <strong>and</strong> intensity<br />

was scored semiquantitatively <strong>and</strong> grouped into cancers with high <strong>and</strong> low expression <strong>of</strong> C4.4A.<br />

Kaplan-Meier survival curves were generated to evaluate the significance <strong>of</strong> C4.4A expression in<br />

prognosis <strong>of</strong> NSCLC patients.<br />

High C4.4A expression was observed in 44% <strong>of</strong> the specimens analyzed, <strong>and</strong> it correlates with<br />

overall survival (P=0.012). Intriguingly, a very strong correlation was noted between high<br />

expression <strong>of</strong> C4.4A in pulmonary adenocarcinoma <strong>and</strong> survival (P


i098i<br />

Urokinase Receptor Splice Variant uPAR-del4/5 <strong>and</strong><br />

rab31 mRNA Expression in Breast Cancer<br />

Magdolen V* 1 , Kotzsch M 2 , Sieuwerts A 3 , Grosser M 2 , Meye A 4 , Smid M 3 , Schmitt M 1 , Luther T 4 ,<br />

Foekens JA 3<br />

1 Institute <strong>of</strong> Pathology, Technical University Dresden, Germany;<br />

2 Clinical Research Unit, Department Obstetrics <strong>and</strong> Gynecology, Technical University München,<br />

Germany;<br />

3 Department <strong>of</strong> Medical Oncology, Erasmus Medical Center Rotterdam, Netherl<strong>and</strong>s;<br />

4 Medizinisches Labor Ostsachsen, Bautzen, Germany<br />

Presenting author e-mail: viktor.magdolen@lrz.tum.de<br />

In the present study, we evaluated the impact <strong>of</strong> mRNA expression <strong>of</strong> the uPAR splice variant<br />

uPAR-del4/5 (lacking exons 4 <strong>and</strong> 5) on metastasis-free survival (MFS) in 280 node-negative<br />

breast cancer patients. Furthermore, microarray techniques were applied to identify differentially<br />

expressed genes associated with high uPAR-del4/5 mRNA levels in tumor tissue <strong>and</strong> tumor cell<br />

lines.<br />

High uPAR-del4/5 mRNA values were strongly associated with shorter MFS <strong>of</strong> breast cancer<br />

patients (P=0.001). In multivariate analysis, uPAR-del4/5 significantly contributed to the base<br />

model <strong>of</strong> traditional prognostic factors for MFS (HR=3.29). By microarray analyses, the gene<br />

encoding rab31, a member <strong>of</strong> the Ras oncogene superfamily, was identified as one <strong>of</strong> seven<br />

genes to be more than 2-fold upregulated in tumor samples <strong>and</strong> cell lines with high uPARdel4/5<br />

mRNA expression. rab31 was selected for analysis <strong>of</strong> mRNA expression in the same set<br />

<strong>of</strong> 280 patients. High rab31 mRNA values were significantly associated with shorter DMFS in<br />

multivariate analysis (HR=2.27, P


i099i<br />

PN-1, a Serine Protease Inhibitor, Increases MMP-9 Activity<br />

in Breast Cancer Cell Line<br />

Fayard B* <strong>and</strong> Monard D<br />

Friedrich Miesher Institute, Basel, Switzerl<strong>and</strong><br />

Presenting author e-mail: berengere.fayard@fmi.ch<br />

Protease Nexin-1 (PN-1) belongs to the serpin family (serine protease inhibitor), acting on a<br />

broad range <strong>of</strong> protease including tPA, uPA, FactorXIa <strong>and</strong> thrombin. PN-1 is overexpressed in<br />

many malignant tumors, its role in cancer remains however unknown. To study its involvement<br />

in tumor cells <strong>and</strong> particularly in invasion, we used four cell lines isolated from the same mouse<br />

mammary tumor named as 67NR, 168FARN, 4T07 <strong>and</strong> 4T1 but displaying different metastatic<br />

properties once reinjected in the mouse mammary fat pad. Here we show that, in these cell lines,<br />

the level <strong>of</strong> Matrix Metalloproteinase-9 (MMP-9) activity, known to be one <strong>of</strong> the key molecules in<br />

invasion, correlates with PN-1 expression. Indeed, the less invasive cell line, 168FARN, presents<br />

the lowest MMP-9 activity but also does not express PN-1. In contrast, the three other cell lines<br />

showing high levels <strong>of</strong> MMP-9 activity express PN-1.<br />

In 168FARN, addition <strong>of</strong> PN-1 leads to an increase <strong>of</strong> secreted MMP-9. This effect is mediated<br />

through its binding to LRP (lipoprotein receptor-related protein) as indicated by the inhibitory<br />

effect <strong>of</strong> a peptide known to antagonize binding <strong>of</strong> PN-1 to LRP <strong>and</strong> by assays performed in LRP–<br />

/– MEF cells. PN-1 does not lead to an increase in MMP-9 mRNA but an increased intracellular<br />

level <strong>of</strong> the metalloprotease is detected when secretion is blocked by Brefeldin.<br />

Unexpectedly, PN-1 showed ability to increase MMP-9 activity in cancer cell lines by a mechanism<br />

dependent <strong>of</strong> its binding to LRP receptor. These results provide evidence for a new function for<br />

PN-1, which as protease inhibitor, regulates the levels <strong>of</strong> secreted MMP-9 proteinas a mean to<br />

affect the cellular invasive behavior <strong>of</strong> tumor cells.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 111


i100i<br />

Cleavage <strong>of</strong> uPAR: Mechanism <strong>and</strong> Prognostic Significance<br />

Høyer-Hansen G*, Almasi CE, Pappot H<br />

Finsen Laboratory, Copenhagen Biocenter, Copenhagen Denmark<br />

Presenting author e-mail: gunilla@finsenlab.dk<br />

On the cell surface uPA cleaves uPAR(I-III) in the linker region between domains I <strong>and</strong> II <strong>of</strong> uPAR.<br />

Physiological concentrations <strong>of</strong> uPA cleave glycolipid-anchored uPAR but not soluble uPAR<br />

(suPAR). This is due to a difference in the conformation <strong>of</strong> the linker region between domains I<br />

<strong>and</strong> II. In media from PMA-stimulated U937 cells intact suPAR, but not cleaved, suPAR(II-III), is<br />

present even though on the cell surface both uPAR(I-III) <strong>and</strong> uPAR(II-III) are present in similar<br />

amounts. Treatment <strong>of</strong> the cells with Pi-PLC releases both uPAR forms from the cell surface.<br />

Biotinylation <strong>of</strong> cell surface proteins verified the localization <strong>of</strong> both uPAR forms, whereas nonbiotinylated<br />

uPAR(I-III) <strong>and</strong> uPAR(II-III) were found in the water-phase indicating intracellular<br />

location. When uPA mediated uPAR cleavage was inhibited by incubation with a neutralizing<br />

anti-uPA antibody, uPAR(II-III) was not found on the cell surface. The intracellular pattern <strong>of</strong><br />

uPAR(I-III) <strong>and</strong> uPAR(II-III) was not affected by inhibiting the cell surface cleavage. However, in<br />

the media from cells grown in the presence <strong>of</strong> the neutralizing anti-uPA antibody, both suPAR(I-<br />

III) <strong>and</strong> suPAR(II-III) were present. This indicates that uPAR(II-III) is shed from the cell surface<br />

when uPA-mediated cleavage <strong>of</strong> uPAR is blocked.<br />

The total amount <strong>of</strong> all uPAR forms in tumor lysates or blood correlates with prognosis in several<br />

forms <strong>of</strong> cancer. However, the amounts <strong>of</strong> the cleaved forms may be directly related to the uPA<br />

activity <strong>and</strong> in lung cancer uPAR(I) is an even stronger prognostic marker than the total amount<br />

<strong>of</strong> all uPAR forms.<br />

112 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i101i<br />

Expression <strong>of</strong> Urokinase Receptor (uPAR) <strong>and</strong><br />

<strong>Plasminogen</strong> Activator Inhibitor-1 (PAI-1) in Human<br />

Colon Cancer <strong>and</strong> their Matched Liver Metastases<br />

Illemann M* 1 , Bird N 2 , Majeed A 2 , Laerum OD 1,3 , Lund LR 1 , Danø K 1 , Nielsen BS 1<br />

1 The Finsen Laboratory, Rigshospitalet, Copenhagen, Denmark;<br />

2 Academic Surgical Unit, University <strong>of</strong> Sheffield, Sheffield, Engl<strong>and</strong>;<br />

3 The Gade Institute, Haukel<strong>and</strong> University Hospital, Bergen, Norway<br />

Presenting author e-mail: millemann@finsenlab.dk<br />

In primary colon adenocarcinomas, expression <strong>of</strong> uPAR <strong>and</strong> PAI-1 is focally upregulated in<br />

the invasive front, primarily in recruited <strong>and</strong> activated inflammatory cells <strong>and</strong> my<strong>of</strong>ibroblasts.<br />

To compare the expression <strong>of</strong> uPAR <strong>and</strong> PAI-1 in primary colon adenocarcinomas with that<br />

in colon cancer liver metastases, we have analyzed matched samples from 14 patients. In the<br />

primary tumors, we found uPAR-immunoreactivity primarily in macrophages <strong>and</strong> PAI-1immunoreactivity<br />

primarily in my<strong>of</strong>ibroblasts located in the invasive front in agreement with<br />

previous findings. The growth pattern <strong>of</strong> colon cancer liver metastases can be divided into<br />

two major groups, those with desmoplastic encapsulation (desmoplastic growth pattern) <strong>and</strong><br />

those without (solid growth pattern). In the current material we had 8 <strong>and</strong> 6 cases, respectively.<br />

In all liver metastases with desmoplastic <strong>and</strong> one with solid growth pattern we found uPARimmunoreactivity<br />

primarily in macrophages <strong>and</strong> neutrophils located in the tumor edge,<br />

whereas the remaining 5 cases showing solid growth had uPAR-immunoreactivity almost only<br />

in neutrophils. PAI-1 was in cases with desmoplastic growth seen in my<strong>of</strong>ibroblasts within<br />

the desmoplastic capsule, while metastases with solid growth had only a few PAI-1 positive<br />

my<strong>of</strong>ibroblasts. Our findings suggest that metastasis-induced desmoplasia with recruitment <strong>of</strong><br />

macrophages upregulate the expression <strong>of</strong> uPAR <strong>and</strong> PAI-1 in a pattern similar to that in the<br />

primary tumors. The metastases with solid growth, in contrast, do not induce uPAR expression in<br />

the recruited macrophages <strong>and</strong> PAI-1 is limited to the sporadically appearing my<strong>of</strong>ibroblasts.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 113


i102i<br />

A Structural Basis for Differential Cell Signaling Initiated by<br />

PAI-1 <strong>and</strong> PAI-2: Implications for Metastatic Potential<br />

Croucher D* 1 , Saunders D 1 , Ranson M 2<br />

1 Cancer Research Program, Garvan Institute <strong>of</strong> Medical Research, Sydney, Australia;<br />

2 School <strong>of</strong> Biological Sciences, University <strong>of</strong> Wollongong, Wollongong, Australia<br />

Presenting author e-mail: d.croucher@garvan.org.au<br />

Tumour expression <strong>of</strong> the uPA inhibitor PAI-1 strongly correlates with poor patient prognosis.<br />

However, tumour expression <strong>of</strong> the related uPA inhibitor PAI-2 generally correlates with good<br />

patient prognosis. Currently, there are no adequate biochemical/functional data to explain this<br />

discrepancy.<br />

One theory for the tumour-promoting effects <strong>of</strong> PAI-1 is based on the signaling events initiated<br />

upon PAI-1 inhibition <strong>of</strong> cell surface uPA. On MCF-7 cells, uPA binding to uPAR induces a<br />

transient pulse <strong>of</strong> ERK activation. Inhibition by PAI-1 unveils a cryptic high affinity site within the<br />

PAI-1 moiety that binds to members <strong>of</strong> the LDLR endocytosis receptor family. Binding <strong>of</strong> this high<br />

affinity site in PAI-1 to VLDLr sustains ERK phosphorylation <strong>and</strong> initiates cell proliferation via an<br />

as yet unknown mechanism.<br />

Through biochemical analysis we have shown that unlike PAI-1, the PAI-2 moiety <strong>of</strong> uPA:PAI-2<br />

does not contain a high affinity site for LDLR members, although the uPA:PAI-2 complex is still<br />

efficiently endocytosed by these receptors. Therefore, upon inhibition <strong>of</strong> uPA, PAI-2 does not<br />

induce the sustained ERK phosphorylation <strong>and</strong> proliferation associated with PAI-1 inhibition.<br />

This differential binding <strong>and</strong> subsequent signal propagation is possibly due to an incomplete<br />

LDLR binding motif within the helix-D <strong>of</strong> PAI-2, which is complete in the helix-D PAI-1.<br />

These data present a possible mechanism by which PAI-2 is able to clear cell surface uPA, <strong>and</strong><br />

hence proteolytic activity, via VLDLr without initiating the cell signaling events <strong>and</strong> increased<br />

metastatic potential associated with high PAI-1. This may partially explain why high tumour<br />

levels <strong>of</strong> PAI-1 correlate with poor prognosis, whereas high tumour levels <strong>of</strong> PAI-2 correlate with<br />

good prognosis.<br />

114 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i103i<br />

Thrombin Induces Tumor Invasion through the Induction<br />

<strong>and</strong> Association <strong>of</strong> Matrix Metalloproteinase-9 <strong>and</strong> b-1 Integrin<br />

on the Cell Surface<br />

Bruno K* 1 , Radjabi R 2 , Sawada K 2 , Montag A 3,4 , Kossiak<strong>of</strong>f A 1 , Lengyel E 2<br />

1 Biochemistry <strong>and</strong> <strong>Molecular</strong> <strong>Biology</strong>, 2 Departments <strong>of</strong> Obstetrics <strong>and</strong> Gynecology/Section <strong>of</strong><br />

Gynecologic Oncology, 3 Pathology, 4 Committee on Cancer <strong>Biology</strong>, University <strong>of</strong> Chicago, Chicago,<br />

Illinois, USA<br />

Presenting author e-mail: kbruno@uchicago.edu<br />

The procoagulatory serine protease thrombin is known to induce invasion <strong>and</strong> metastasis in<br />

various cancers. However, the mechanisms by which it promotes tumorigenesis are poorly<br />

understood. Since the 92 kDa gelatinase (MMP-9) is a known mediator <strong>of</strong> tumor cell invasion,<br />

we sought to determine if <strong>and</strong> how thrombin regulates MMP-9. The thrombin receptor,PAR-1, is<br />

expressed in osteosarcomas, as determined by immunohistochemistry <strong>and</strong> RT-PCR. Stimulation<br />

<strong>of</strong> U2-OS osteosarcoma cells with thrombin <strong>and</strong> a thrombin receptor activating peptide (TRAP)<br />

induced pro-MMP-9 secretion as well as cell surface associated pro-MMP-9 expression <strong>and</strong><br />

proteolytic activity. This was paralleled by an increase in MMP-9 mRNA <strong>and</strong> MMP-9 promoter<br />

activity. Thrombin induced invasion <strong>of</strong> U2-OS cells through matrigel could be inhibited with<br />

a MMP-9 antibody <strong>and</strong> was mediated by the PI3-kinase signaling pathway. The stimulation<br />

<strong>of</strong> MMP-9 by thrombin was paralleled by an increase in b1-integrin mRNA <strong>and</strong> b1-integrin<br />

expression on the cell surface, which was also mediated by PI3-kinase <strong>and</strong> was also required for<br />

invasion. Thrombin activation induced <strong>and</strong> co-localized both b1-integrin <strong>and</strong> pro-MMP-9 on the<br />

cell membrane, as evidenced by coprecipitation <strong>and</strong> confocal microscopy. The thrombin mediated<br />

association <strong>of</strong> these two proteins, as well as thrombin mediated invasion <strong>of</strong> U2-OS cells could be<br />

blocked with a cyclic peptide <strong>and</strong> with an antibody preventing binding <strong>of</strong> the MMP-9 hemopexin<br />

domain to the b1-integrin I-like domain. These results suggest that thrombin induces expression<br />

<strong>and</strong> association <strong>of</strong> b1-integrin with MMP-9 <strong>and</strong> that the cell surface localization <strong>of</strong> the protease by<br />

the integrin promotes tumor cell invasion.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 115


i104i<br />

Overexpression <strong>of</strong> Protease Nexin-1 mRNA in<br />

Oral Squamous Cell Carcinomas<br />

Gao S 1 , Krogdahl A 2 , Sørensen JA 3 , Dabelsteen E 4 , Andreasen PA* 1<br />

1 Department <strong>of</strong> <strong>Molecular</strong> <strong>Biology</strong>, University <strong>of</strong> Aarhus, Aarhus, Denmark;<br />

2 Department <strong>of</strong> Pathology, Odense University Hospital, Odense, Denmark;<br />

3 Department <strong>of</strong> Plastic Surgery, Odense University Hospital, Odense, Denmark;<br />

4 School <strong>of</strong> Dentistry, University <strong>of</strong> Copenhagen, Copenhagen, Denmark<br />

Presenting author e-mail: pa@mb.au.dk<br />

Protease nexin-1 (PN-1) belongs to the serpin family <strong>of</strong> serine protease inhibitors. It is the<br />

phylogenetically closest relative <strong>of</strong> plasminogen activator inhibitor-1 (PAI-1). PN-1 is less specific<br />

than PAI-1, inhibiting, besides uPA, also plasmin <strong>and</strong> thrombin at physiologically relevant rates.<br />

While there are numerous studies <strong>of</strong> the occurrence <strong>and</strong> functions <strong>of</strong> PAI-1 in cancer, a possible<br />

tumour biological role <strong>of</strong> PN-1 has been almost totally neglected. We have now initiated studies<br />

<strong>of</strong> the occurrence <strong>of</strong> PN-1 in tumours, also with the presumption that investigations <strong>of</strong> the tumour<br />

biological functions <strong>of</strong> PN-1 may provide clues to the uncertainty about the molecular <strong>and</strong><br />

cellular mechanisms underlying the correlation between a high PAI-1 level in tumours <strong>and</strong> a poor<br />

prognosis for the patient. We compared the level <strong>of</strong> PN-1 mRNA in 20 cases <strong>of</strong> oral squamous cell<br />

carcinomas <strong>and</strong> in matched samples <strong>of</strong> the corresponding normal oral tissues. We found that the<br />

average PN-1 levels in tumours <strong>and</strong> normal tissues were significantly different, being increased<br />

up to 13 fold in tumour samples compared with the average level in normal tissues. The PN-1<br />

mRNA level was significantly higher in tumours from patients with lymph node metastasis than<br />

in tumours from patients without. We could conclude that PN-1 is frequently overexpressed in<br />

oral squamous cell carcinoma <strong>and</strong> may correlate with lymph node metastasis. Development <strong>of</strong> an<br />

anti-PN-1 antibody suitable for immunohistochemical identification <strong>of</strong> PN-1 expressing cell types<br />

in different carcinoma types may help us to underst<strong>and</strong> its tumour biological functions.<br />

116 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i105i<br />

The Matrix Metalloprotease (MMP) Inhibitor Galardin Increases<br />

Collagen Deposition <strong>and</strong> Reduces Spontaneous Metastasis<br />

in the MMTV-PymT Transgenic Breast Cancer Model<br />

Almholt K* 1 , Lærum OD 2 , Lund LR 1 , Danø K 1 , Johnsen M 3 , Rømer J 1<br />

1 Finsen Laboratory, Rigshospitalet, Copenhagen, Denmark;<br />

2 The Gade Institute, Section <strong>of</strong> Pathology, Haukel<strong>and</strong> University Hospital, Bergen, Norway;<br />

3 Department <strong>of</strong> <strong>Molecular</strong> <strong>Biology</strong>, University <strong>of</strong> Copenhagen, Copenhagen, Denmark<br />

Presenting author e-mail: kasper@finsenlab.dk<br />

Matrix metalloproteases (MMPs) <strong>and</strong> other extracellular proteases are recognized as critical<br />

factors in cancer progression. MMP inhibitors (MMPIs) have proven effective primarily by<br />

reducing growth <strong>of</strong> xenograft tumors in mice. However, small molecule MMPIs have not been<br />

tested in any transgenic/spontaneous tumor model. We chose the MMTV-PymT transgenic breast<br />

cancer model to analyze the effect <strong>of</strong> galardin/GM6001, a potent MMPI that reacts with most<br />

MMPs. Prior to the experiment on cancer-bearing mice we determined a treatment regimen based<br />

on the ability to inhibit skin wound healing in mice. We then followed a cohort <strong>of</strong> 54 MMTV-<br />

PymT-transgenic mice that received either galardin or placebo. Treatment was started at age 6<br />

weeks with galardin at 100 mg/kg/day in the form <strong>of</strong> s.c. implanted slow-release tablets. The<br />

growth <strong>of</strong> the primary tumors was reduced but not severely retarded by galardin treatment. All<br />

mice were killed at age 13½ weeks, at which time the average primary tumor burden was reduced<br />

to 1.69 cm3 in galardin-treated mice compared to 3.29 cm3 in control mice (t-test, p=0.0014). It is<br />

likely that some MMPs promote while others inhibit tumor growth, <strong>and</strong> the net effect <strong>of</strong> inhibiting<br />

a whole range <strong>of</strong> MMPs was not sufficient to prevent tumor growth. However, the tumors<br />

from galardin-treated mice <strong>of</strong>ten had a lower tumor grade as determined by histopathological<br />

evaluation. The tumors from galardin-treated mice also showed a much higher degree <strong>of</strong> collagen<br />

deposition compared to tumors from placebo-treated mice. We then quantified the total lung<br />

metastases volume in the same mice. The median metastasis volume was reduced to 0.003 mm3 in<br />

galardin-treated mice compared to 0.56 mm3 in control mice (t-test, p


i106i<br />

Proteomics <strong>of</strong> uPAR Protein: Protein Interactions in Cancer Metastasis<br />

Saldanha R, Molloy M, Xu N, Baker MS*<br />

Australian Proteome Analysis Facility Ltd <strong>and</strong> CORE in Biomolecular Frontiers, Macquarie University,<br />

Sydney, Australia<br />

Presenting author e-mail: mbaker@proteome.org.au<br />

uPAR is a key player in cancer metastasis. Our past studies show that when metastatic HCT116<br />

cells express a 5’ uPAR antisense cDNA fragment this results in (i) reduced uPAR cell surface<br />

expression; (ii) reduced cell-surface uPA binding, (iii) reduced adhesion <strong>and</strong> plasminogendependant<br />

matrix degradation, (iv) reduced Erk MAP kinase activity, Src kinase activity <strong>and</strong><br />

MMP-9 secretion; (v)reduced metastasis in Nu/Nu mice <strong>and</strong> (vi) new data that shows statistically<br />

significant expression/phosphorylation changes in only a small percentage (~4%) <strong>of</strong> the ~850<br />

proteins reproducibly detected by 2DE (e.g., stathmin, C-Myc binding protein, translation<br />

initiation factor 3, tropomyosin 3, histone H2B type 12, pr<strong>of</strong>ilin-2, importin, HSP90 to name but<br />

a few). To further characterize how uPAR operates we have now identified (<strong>and</strong> are quantifying<br />

using IP-iTRAQ) uPAR protein binding complexes from malignant OVCA 429 cells. Proteins<br />

that specifically co-purified with anti-uPAR mAbs were validated using orthogonal approaches<br />

(e.g., cross-over IP Westerns). Of particular interest were a range <strong>of</strong> signal transduction proteins,<br />

proteins previously associated with plasminogen activation (phospholipase C, LRP, <strong>and</strong> annexin<br />

A2, enolase, thrombospondin) <strong>and</strong> some novel proteins (TGF-bR2 <strong>and</strong> only the integrin, av-b6<br />

(although these cells express other integrins). Physical sites <strong>of</strong> interaction between uPAR <strong>and</strong> avb6<br />

were definitively identified by peptide mapping strategies as was direct interaction <strong>of</strong> TGF-bR2<br />

with uPAR. Functional consequences <strong>of</strong> uPAR interactions are being studied <strong>and</strong> preliminary data<br />

show that uPAR mediates many biological effects through protein:protein interactions.<br />

118 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i107i<br />

A New Tagging System for Production <strong>of</strong> Recombinant Proteins in<br />

Drosophila S2 Cells Using the Third Domain <strong>of</strong> the Urokinase Receptor<br />

Gårdsvoll H* 1 , Hansen LV 1 , Jørgensen TJD 2 , Ploug M 1<br />

1 Finsen Laboratory, Rigshospitalet, Copenhagen, Denmark;<br />

2 Department <strong>of</strong> Biochemistry <strong>and</strong> <strong>Molecular</strong> <strong>Biology</strong>, University <strong>of</strong> Southern Denmark, Odense,<br />

Denmark<br />

Presenting author e-mail: gvoll@finsenlab.dk<br />

The use <strong>of</strong> protein fusion tag technology greatly facilitates detection, expression <strong>and</strong> purification<br />

<strong>of</strong> recombinant proteins, <strong>and</strong> the dem<strong>and</strong>s for new <strong>and</strong> more effective systems are therefore<br />

exp<strong>and</strong>ing. We have used a soluble truncated form <strong>of</strong> the third domain <strong>of</strong> the urokinase<br />

receptor as a convenient C-terminal fusion partner for various recombinant extracellular human<br />

proteins used in basic cancer research. The stability <strong>of</strong> this cystein-rich domain, which structure<br />

adopts a three-finger fold, provides an important asset for its applicability as a fusion tag for<br />

expression <strong>of</strong> recombinant proteins. Up to 20 mg <strong>of</strong> intact fusion protein were expressed by<br />

stably transfected Drosophila S2 cells per liter <strong>of</strong> culture using this strategy. Purification <strong>of</strong> these<br />

secreted fusion proteins from the conditioned serum free medium <strong>of</strong> S2 cells was accompanied<br />

by an efficient one-step immunoaffinity chromatography procedure using the immobilized<br />

anti-uPAR monoclonal antibody R2. An optional enterokinase cleavage site is included between<br />

the various recombinant proteins <strong>and</strong> the linker region <strong>of</strong> the tag, which enables generation <strong>of</strong><br />

highly pure preparations <strong>of</strong> tag-free recombinant proteins. Using this system we successfully<br />

produced soluble <strong>and</strong> intact recombinant forms <strong>of</strong> extracellular proteins such as CD59, C4.4A <strong>and</strong><br />

vitronectin, as well as a number <strong>of</strong> truncated domain constructs <strong>of</strong> these proteins. In conclusion,<br />

the present tagging system <strong>of</strong>fers a convenient general method for the robust expression <strong>and</strong><br />

efficient purification <strong>of</strong> a variety <strong>of</strong> recombinant proteins.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 119


i108i<br />

Photoaffinity Labeling <strong>of</strong> uPAR with Cyclic Peptides<br />

Jacobsen B* 1 , Gårdsvoll H 1 , Barkholt V 2 , Østergaard S 3 , Ploug M 1<br />

1 Finsen Laboratory, Rigshospitalet, Copenhagen Biocenter, Denmark;<br />

2 Enzyme <strong>and</strong> Protein Chemistry, BioCentrum-DTU, The Technical University <strong>of</strong> Denmark, Denmark;<br />

3 Novo Nordisk A/S, Novo Research Park, Denmark<br />

Presenting author e-mail: bjacobsen@finsenlab.dk<br />

In view <strong>of</strong> its involvement in cancer invasion <strong>and</strong> metastasis, the high-affinity uPA-uPAR<br />

interaction represents an attractive target for rational drug design, thus requiring detailed<br />

knowledge <strong>of</strong> the receptor-lig<strong>and</strong> interface. In this study, we use photoaffinity labeling as a tool<br />

for delineation <strong>of</strong> structural aspects <strong>of</strong> this interaction. The technique enables the covalent crosslinking<br />

<strong>of</strong> two molecules by excitation <strong>of</strong> a photoactivatable probe, which thereby can abstract a<br />

hydrogen atom from a C-H-bond in its vicinity. This has been used successfully to identify the<br />

uPAR-binding sites <strong>of</strong> linear phage-display peptides. We have now synthesized a range <strong>of</strong> cyclic<br />

peptides mimicking the long b-hairpin <strong>of</strong> the growth factor-like domain <strong>of</strong> uPA, which includes<br />

residues essential for its binding to uPAR (Tyr24, Phe25, Ile28 <strong>and</strong> Trp30). In the peptides AE03<br />

<strong>and</strong> AE04, the photoprobe p-benzoyl-L-phenylalanine has been inserted at positions equivalent to<br />

Trp30 <strong>and</strong> Tyr24 in uPA, respectively, <strong>and</strong> biotin has been included as a detection <strong>and</strong> purification<br />

tag. Photolysis <strong>of</strong> uPAR-peptide complexes causes a covalent incorporation <strong>of</strong> the cyclic peptides<br />

into uPAR. The specificity <strong>of</strong> this insertion reaction was certified by the inhibition obtained when<br />

preincubating with receptor-binding derivatives <strong>of</strong> uPA. By cleavage with chymotrypsin <strong>and</strong><br />

western blotting with streptavidin, AE03 is shown to incorporate into uPAR domain I, while AE04<br />

is domains II + III-specific. These results are supported by the recently solved crystal structure<br />

<strong>of</strong> uPAR in complex with the amino-terminal fragment <strong>of</strong> uPA, revealing a hydrophobic lig<strong>and</strong>binding<br />

cavity formed by the assembly <strong>of</strong> all three domains <strong>of</strong> uPAR.<br />

120 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i109i<br />

In Vivo Inhibition <strong>of</strong> the Murine uPA-uPAR Interaction using<br />

Monoclonal Antibodies Raised in uPAR Deficient Mice<br />

Rasch MG*, Pass J, Jögi A, Rønø B, Gårdsvoll H, Lund LR, Høyer-Hansen G, Lund IK<br />

Finsen Laboratory, Copenhagen Biocenter, Copenhagen, Denmark<br />

Presenting author e-mail: mrasch@finsenlab.dk<br />

The interaction between uPA <strong>and</strong> uPAR is a potential target for anti-invasive cancer therapy.<br />

Cancer invasion <strong>and</strong> metastasis is dependent on the interaction between cancer <strong>and</strong> stromal cells,<br />

which can be studied in genetically induced murine cancer models. To enable in vivo studies <strong>of</strong><br />

the effect on inhibition <strong>of</strong> uPA-uPAR binding in genetically induced murine cancer models, we<br />

have generated murine mAbs against murine uPAR (muPAR) by immunizing uPAR–/– mice with<br />

recombinant muPAR.<br />

We have selected five mAbs (mR1-mR5) <strong>of</strong> which mR1, mR3, <strong>and</strong> mR5 recognize epitopes in<br />

domain I <strong>of</strong> muPAR, while mR2 <strong>and</strong> mR4 recognize epitopes in domain II-III <strong>of</strong> muPAR. In cell<br />

binding experiments mR1 <strong>and</strong> mR4 antagonized the binding <strong>of</strong> 125I-mATF with IC50 values <strong>of</strong><br />

0.67 nM for mR1 <strong>and</strong> 0.40 nM for mR4 compared to 0.14 nM for unlabeled mATF. Additionally,<br />

mR1 could rescue 50% P388D.1 cells in an anthrax-toxin based assay, which requires cell-bound<br />

uPA to kill the cells. In contrast to all previously generated uPAR specific mAbs, mR3 binds<br />

domain I without interfering with uPA binding.<br />

In vivo efficacy <strong>of</strong> mR1 was demonstrated by the ability <strong>of</strong> mR1 to rescue mice treated with a lethal<br />

dose <strong>of</strong> uPA-activatable anthrax toxins. Another in vivo approach showed that mR1 treated<br />

tPA–/– mice mimicked the hepatic fibrin deposits <strong>of</strong> uPAR–/–/tPA–/– mice. Thus, mR1 is<br />

an efficient in vivo antagonist <strong>of</strong> the uPA-uPAR interaction <strong>and</strong> well suited for therapeutic<br />

intervention.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 121


i110i<br />

RNA Interference for Urokinase-Targeting Limits Growth<br />

<strong>of</strong> Hepatocellular Carcinoma Xenografts in Nude Mice<br />

Salvi A*, Arici B, Barlati S, De Petro G<br />

Division <strong>of</strong> <strong>Biology</strong> <strong>and</strong> Genetics, Department <strong>of</strong> Biomedical Sciences <strong>and</strong> Biotechnology, IDET Centre<br />

<strong>of</strong> Excellence, University <strong>of</strong> Brescia, Brescia, Italy<br />

Presenting author e-mail: asalvi@med.unibs.it<br />

Gene expression <strong>of</strong> urokinase-type plasminogen activator (u-PA) is up-regulated in human<br />

hepatocellular carcinoma (HCC) <strong>and</strong> the levels <strong>of</strong> u-PA mRNA are inversely related to HCC<br />

patients’ survival. The purpose <strong>of</strong> this study was to examine the effects <strong>of</strong> vector-based RNA<br />

interference (RNAi) <strong>of</strong> u-PA on the growth <strong>of</strong> human HCC xenografts in nude mice in order to<br />

investigate the u-PA role in human HCC.<br />

Our results showed that the s.c. injection <strong>of</strong> siRNA u-PA SKHep1C3 stable transfected cells<br />

(pSsiRNAu-PA) led to a growth delay in xenograft development. The molecular characterization<br />

<strong>of</strong> nodules (carried out by PCR, RT-PCR/Agilent technology <strong>and</strong> immunoistochemical analysis)<br />

revealed the presence <strong>of</strong> plasmid DNA, the u-PA gene expression knock-down, at both mRNA<br />

<strong>and</strong> protein levels, giving evidences <strong>of</strong> a long-term <strong>and</strong> target-specific inhibition by vectorbased<br />

RNAi 11 weeks after cell inoculation. Furthermore the immunohistochemical <strong>and</strong><br />

immun<strong>of</strong>luorescence evaluation <strong>of</strong> fibronectin (FN) expression <strong>and</strong> organization revealed FN<br />

fibrils in pSsiRNAu-PA xenografts <strong>and</strong> in pSsiRNA u-PA-cells. These results represent the first<br />

experimental approach to inhibit u-PA in HCC xenografts using an ablative strategy, supporting<br />

the notion that u-PA is involved in HCC development, with a major role for u-PA in the initial<br />

phases <strong>of</strong> HCC growth; providing evidence that u-PA siRNA construct can be maintained in each<br />

xenograft for up to 11 weeks <strong>of</strong> tumour development; also identifying the ability to organize FN<br />

fibrils by SKHep1C3 cells as a downstream effect <strong>of</strong> u-PA knock-down. These data could open<br />

the possibility for experimental HCC gene therapy by vector-based RNAi against u-PA in animal<br />

models.<br />

122 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i111i<br />

Potent <strong>and</strong> Broad Anti-tumor Activity <strong>of</strong> an Engineered<br />

Matrix Metalloproteinase-activated Anthrax Lethal<br />

Toxin that Targets Tumor Vasculature<br />

Liu S 1 , Wang H 1 , Currie BM 2 , Molinolo A 2 , Leung HJ 1 , Moayeri M 1 , Alfano RW 3 , Frankel AE 3 ,<br />

Leppla SH 1 , Bugge TH* 2<br />

1 Laboratory <strong>of</strong> Bacterial Diseases, NIAID, NIH, Bethesda, Maryl<strong>and</strong>, USA;<br />

2 Oral <strong>and</strong> Pharyngeal Cancer Branch, NIDCR, NIH, Bethesda, Maryl<strong>and</strong>, USA;<br />

3 Cancer Research Institute <strong>of</strong> Scott <strong>and</strong> White Memorial Hospital, Temple, Texas, USA<br />

Presenting author e-mail: thomas.bugge@nih.gov<br />

Anthrax lethal toxin (LT), which specifically inactivates MEKs, can treat human melanomas with<br />

BRAF V600E mutations, because <strong>of</strong> their addiction to the MEK-ERK pathway. To more selectively<br />

target tumors, we have developed a matrix-metalloproteinase (MMP)-activated anthrax LT, which<br />

specifically targets tumors overexpressing MMPs. Surprisingly, the MMP-activated LT has potent<br />

anti-tumor activity against a wide range <strong>of</strong> other tumor types, regardless <strong>of</strong> their BRAF mutation<br />

status. Moreover, the engineered toxin could be curative to established malignant solid tumors<br />

when used systemically, despite having very low tissue toxicity. This is largely due to the indirect<br />

targeting <strong>of</strong> tumor angiogenesis. Moreover, the engineered toxin not only exhibits much lower<br />

toxicity than wild-type LT to mice, but also shows higher toxicity to tumors because <strong>of</strong> its greater<br />

bioavailability. Thus, in theory, all tumor types are expected to respond to the MMP-activated LT<br />

therapy, <strong>and</strong> patients with tumors containing the BRAF mutation may derive additional benefits<br />

due to the direct toxicity <strong>of</strong> the toxin to the cancer cells. We are currently preparing a clinical trial<br />

to test the MMP-activated LT in cancer patients.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 123


i112i<br />

Elucidation <strong>of</strong> the Epitope <strong>of</strong> MA-31C9, a Non-inhibitory<br />

Anti-human PAI-1 Antibody<br />

Meissenheimer LM*, Dewilde M, Compernolle G, Declerck PJ, Gils A<br />

Laboratory for Pharmaceutical <strong>Biology</strong>, Katholieke Universiteit Leuven, Leuven, Belgium<br />

Presenting author e-mail: lester.meissenheimer@pharm.kuleuven.be<br />

<strong>Plasminogen</strong> Activator Inhibitor 1 (PAI-1) is the most important physiological inhibitor <strong>of</strong><br />

plasminogen activators in vivo. MA-31C9 is a non-inhibitory monoclonal antibody raised towards<br />

human PAI-1, which is <strong>of</strong>ten used as a control antibody in studies on the evaluation <strong>of</strong> PAI-1<br />

inhibitory antibodies. The goal <strong>of</strong> this study was to elucidate the epitope <strong>of</strong> MA-31C9.<br />

Wild-type human PAI-1 has a high affinity for MA-31C9 (i.e. KA=6.8±1.9x109M-1 using surface<br />

plasmon resonance (SPR)) whereas rat PAI-1 does not bind to MA-31C9 (KA


i113i<br />

Residues outside the Epitope Determine the Function <strong>of</strong><br />

MA-159M12, an Inhibitory Anti-rat PAI-1 Antibody<br />

Meissenheimer LM*, Compernolle G, Declerck PJ, Gils A<br />

Laboratory for Pharmaceutical <strong>Biology</strong>, Katholieke Universiteit Leuven, Leuven, Belgium<br />

Presenting author e-mail: lester.meissenheimer@pharm.kuleuven.be<br />

Rat models have been shown to be suitable for in vivo investigations on thrombolysis <strong>and</strong><br />

fibrinolysis. MA-159M12 is a monoclonal antibody raised towards rat PAI-1 that exerts an<br />

inhibitory effect by accelerating the active to latent conversion. The goal <strong>of</strong> this study was to study<br />

the mechanism <strong>of</strong> action <strong>of</strong> MA-159M12.<br />

MA-159M12 has a high affinity for rat PAI-1 (KA=3.4±1.1x109M-1) but does not bind to human<br />

PAI-1 (KA


i114i<br />

Conformational Probes <strong>and</strong> Activity Regulators <strong>of</strong> <strong>Plasminogen</strong><br />

Activator Inhibitor-1, Isolated from Phage-displayed<br />

Disulphide Bridge-constrained Peptide Libraries<br />

Dupont DM, Jensen JK, Mathiasen L, Blouse GE, Wind T, Andreasen PA*<br />

Department <strong>of</strong> <strong>Molecular</strong> <strong>Biology</strong>, University <strong>of</strong> Aarhus, Aarhus, Denmark<br />

Presenting author e-mail: pa@mb.au.dk<br />

To find new probes <strong>and</strong> inhibitors <strong>of</strong> PAI-1s molecular interactions <strong>and</strong> conformational<br />

changes, we screened a phage-displayed peptide library containing 109 different sequences<br />

with the formats X7, CX7C, CX10C, <strong>and</strong> CX3CX3CX3C with human PAI-1 as bait. We isolated<br />

phages harbouring three families <strong>of</strong> peptides. One family <strong>of</strong> peptides had a consensus core<br />

motif <strong>of</strong> CFGaC, referred to a paionin-1, in which a is an aromatic residue. As determined by<br />

site-directed mutagenesis, paionin-1 bound to in a hydrophobic pocket under a-helix D <strong>and</strong><br />

inhibited the binding <strong>of</strong> uPA-PAI-1 complex to the endocytosis receptors LRP <strong>and</strong> VLDLR. A<br />

second family <strong>of</strong> peptide sequences with the formats CX7C <strong>and</strong> CX10C <strong>and</strong> the common core<br />

sequence WPRY bound to the area <strong>of</strong> PAI-1 becoming exposed upon detachment <strong>of</strong> b-str<strong>and</strong> 1C<br />

during latency transition. The access <strong>of</strong> paionin-3 to its binding site was blocked by the glycans<br />

attached to Asn267 in PAI-1 expressed in mammalian cells. A third peptide was <strong>of</strong> the format<br />

CX3CX3CX3CX5CX3CX3CX3C, will be referred to as paionin-4, <strong>and</strong> probably resulted from<br />

an unplanned cloning artefact in the library. Paionin-4 accelerated the rate <strong>of</strong> conversion <strong>of</strong><br />

PAI-1 to the latent state <strong>and</strong> had a binding area on PAI-1 overlapping with that <strong>of</strong> the latencyinducing<br />

monoclonal antibody MA33B8. Binding analyses showed that paionin-4 has a stronger<br />

affinity to conformational forms <strong>of</strong> PAI-1 with an inserted RCL than to active PAI-1. Thus, some<br />

<strong>of</strong> the isolated peptides are conformational probes, while others represent novel approaches to<br />

pharmacological interference with pathophysiological functions <strong>of</strong> PAI-1.<br />

126 X I t h I n t e r n a t i o n a l W o r k s h o p o n


i115i<br />

Urokinase-type <strong>Plasminogen</strong> Activator-inhibiting Cyclic Peptides<br />

Demonstrate New Modalities for Inhibition <strong>of</strong> Serine Proteases<br />

Andersen LM* 1 , Wind T 1 , Hansen HD 1 , Blouse GE 1 , Christensen A 1 , Jensen JK 1 , Malmendal A 2 ,<br />

Nielsen NC 2 , Andreasen PA 1<br />

1 Department <strong>of</strong> <strong>Molecular</strong> <strong>Biology</strong>, 2 Department <strong>of</strong> Chemistry, University <strong>of</strong> Aarhus, Aarhus, Denmark<br />

Presenting author e-mail: lma@mb.au.dk<br />

In order to find new principles for inhibition <strong>of</strong> the enzymatic activity <strong>of</strong> urokinase-type<br />

plasminogen activator (uPA), we screened phage-displayed r<strong>and</strong>om, disulphide bridgeconstrained<br />

peptide repertoires with human <strong>and</strong> murine uPA as baits. With human uPA, the most<br />

frequently isolated sequence was CSWRGLENHRMC, referred to as upain-1. With murine uPA,<br />

the most frequent isolated sequence was CPAYSRYLDC, referred to as mupain-1. The selected<br />

peptide sequences inhibited human or murine uPA competitively, with Ki values around 500 nM.<br />

The Ki values were in excellent agreement with the KD values determined by BIACORE analysis.<br />

By an inhibitory screen against other murine <strong>and</strong> human serine proteases, including trypsin, both<br />

upain-1 <strong>and</strong> mupain-1 were found to be highly selective for human <strong>and</strong> murine uPA, respectively.<br />

However, upain-1 did not measurably inhibit murine uPA <strong>and</strong> mupain-1 did not measurably<br />

inhibit human uPA. Site-directed mutagenesis <strong>of</strong> the peptides as well as <strong>of</strong> the enzymes identified<br />

Arg4 <strong>of</strong> upain-1 <strong>and</strong> Arg6 <strong>of</strong> mupain-1 as the P1 residues <strong>and</strong> indicated that binding specificity<br />

depends on extended binding interactions involving specific surface loops <strong>of</strong> the enzymes <strong>and</strong><br />

several residues <strong>of</strong> the peptides. Comparison <strong>of</strong> the solution structure <strong>of</strong> upain-1, as determined<br />

by NMR, <strong>and</strong> the results <strong>of</strong> the site-directed mutagenesis indicated that conformational changes <strong>of</strong><br />

the peptide are a prerequisite for upain-1-uPA binding. Peptide-derived inhibitors such as upain-<br />

1 <strong>and</strong> mupain-1 provide novel mechanistic information about enzyme-inhibitor <strong>and</strong> enzymesubstrate<br />

interactions, important tools for dissecting structural determinants <strong>of</strong> substrate <strong>and</strong><br />

inhibitor recognition, <strong>and</strong> alternative methodologies for designing effective protease inhibitors.<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 127


i116i<br />

In vivo Treatment with Monoclonal Antibodies against Mouse<br />

Urokinase-type <strong>Plasminogen</strong> Activator in Cancer Models<br />

Jögi A*, Lund IK, Høyer-Hansen G, Lund LR, Danø K, Rømer J<br />

Finsen Laboratory, Rigshospitalet, Copenhagen Biocenter, København, Denmark<br />

Presenting author e-mail: ajogi@finsenlab.dk<br />

Degradation <strong>of</strong> extra cellular matrix is pivotal to tumor metastasis <strong>and</strong> invasive growth <strong>and</strong><br />

plasmin has a well-documented role in both processes. In a transgenic breast cancer model uPAdeficiency<br />

was shown to confer reduced frequency <strong>of</strong> metastasis (Almholt et al (2005) Int J Cancer).<br />

Inhibition <strong>of</strong> the activity <strong>of</strong> uPA is hence a potential approach to anti-invasive cancer therapy. In<br />

order to evaluate this strategy preclinically in mice, reagents are needed that inhibit the activity <strong>of</strong><br />

murine uPA. This protein <strong>and</strong> its associated factors is highly species specific <strong>and</strong> although several<br />

strong inhibitors <strong>of</strong> human uPA activity are known, none <strong>of</strong> these are efficient inhibitors in the<br />

mouse. In order to develop such inhibitors, we have recently, by immunization <strong>of</strong> uPA-deficient<br />

mice with mouse uPA, generated anti-catalytic monoclonal antibodies (mAbs) against murine<br />

uPA. Here we report the first biological effects from systemic administration <strong>of</strong> such an inhibitory<br />

anti-mouse uPA mAb. In a wound healing model, where migrating leading edge keratinocytes<br />

model the invasive tumor cells, i.p. injections <strong>of</strong> anti-uPA mAb in tPA-deficient mice shifted the<br />

healing time toward that seen in tPA;uPA double-deficient mice. Increasing dose <strong>of</strong> the mAb led<br />

to increasingly delayed healing time in a dose dependent manner, approaching that <strong>of</strong> the genedeficient<br />

mice. Immunoblotting <strong>of</strong> protein extracts from the wounds reveal that less plasmin is<br />

generated in wounds <strong>of</strong> anti-uPA treated mice, compared to none-treated or mock treated mice.<br />

We are currently testing this mAb in a transgenic breast cancer model.<br />

128 X I t h I n t e r n a t i o n a l W o r k s h o p o n


Author Index<br />

(The authors present at the meeting are indicated in bold characters.)<br />

Aguilar, S, 073<br />

Ahlskog, N, 093<br />

Alfano, D, 041, 087, 111<br />

Alfano, RW, 111<br />

Allen, BJ, 072<br />

Almasi, CE, 100<br />

Almholt, K, 032, 096, 105<br />

Ampurdanés, C, 073<br />

Andersen, LM, 115<br />

Andolfo, A, 003, 051, 052<br />

Andreasen, PA, 004, 039, 048, 049,<br />

050, 053, 054, 071, 104, 114, 115<br />

Andreé, J, 010<br />

Antalis, TM, 079<br />

Ardite, E, 040<br />

Arici, B, 110<br />

Arnaudova, R, 068<br />

Astrup, A, 060<br />

Bae Kim, G, 046<br />

Baker, MS, 105<br />

Balsara, RD, 028, 086<br />

Barkholt, V, 108<br />

Barlati, S, 110<br />

Battey, F, 008<br />

Beaufort, N, 037<br />

Bednarek, R, 075<br />

Behrendt, N, 033, 034, 047<br />

Belorgey, D, 023, 091<br />

Betsholtz, C, 010<br />

Bian, C, 048<br />

Bifulco, K, 062<br />

Binder, BR, 031, 044, 061, 063<br />

Bird, N, 101<br />

Blasi, F, 035, 068<br />

Blomstr<strong>and</strong>, F, 067<br />

Blouse, GE, 004, 049, 053, 054, 114,<br />

115<br />

Bødker, JS, 053<br />

Boncela, J, 075<br />

Borger Rasmussen, H, 078<br />

Bøtkjær, KA, 049<br />

Brellier, F, 018<br />

Breuss, JM, 044<br />

Brotzge, XH, 006<br />

Brumwell, A, 013<br />

Brunner, PM, 031, 063<br />

Brünner, N, 027, 060, 084<br />

Bruno, K, 103<br />

Bugge, TH, 024, 025, 032, 033, 034,<br />

046, 078, 079, 111<br />

Byszuk, O, 049<br />

Cale, J, 010<br />

Cantelmo, AR, 062<br />

Caputi, M, 074<br />

Carlsson, L, 083<br />

Carriero, MV, 026, 041, 062<br />

Carroll, VA, 044<br />

Casslén, B, 070<br />

Castellino, FJ, 028, 038, 086<br />

Chapman, HA, 013, 042<br />

Chen, J, 045<br />

Christensen, A, 115<br />

Church, S, 009<br />

Cierniewski, CS, 075, 076<br />

Cochran, BJ, 059<br />

Cohen-Gould, L, 009<br />

Compernolle, G, 112, 113<br />

Corominas, JM, 073<br />

Croucher, D, 058, 059, 080, 102<br />

Cunningham, O, 003<br />

Currie, B, 025, 079, 111<br />

Dabelsteen, E, 071, 104<br />

Dahl, L, 083<br />

D’Alessio, S, 035<br />

Danø, K, 043, 095, 101, 105, 116<br />

De Petro, G, 110<br />

Declerck, PJ, 112, 113<br />

Degryse, B, 068, 069<br />

Del Rosso, M, 065<br />

Demyanets, S, 094<br />

Deora, AB, 009<br />

Deryugina, EI, 049<br />

Devy, L, 045<br />

Dewilde, M, 112<br />

Di Carluccio, G, 062<br />

Diaz-Ramos, A, 082<br />

Dransfield, DT, 045<br />

Dumler, I, 085, 090<br />

Dunoyer-Geindre, S, 012<br />

Dupont, DM, 050, 053, 054, 114<br />

Eden, G, 068<br />

Edwards, DR, 030<br />

Ehart, M, 061<br />

Ehnman, M, 015<br />

Ellis, V, 030<br />

Engelholm, LH, 033, 034<br />

Eriksson, U, 010, 015<br />

Espitia, C, 092<br />

Estrada, Y, 041<br />

Fayard, B, 099<br />

Fibbi, G, 065<br />

Fischer, C, 020<br />

Fish, RJ, 012<br />

Flint, A, 060<br />

Flugel, D, 011<br />

Foekens, JA, 098<br />

Folestad, E, 010<br />

Franco, P, 041, 074<br />

Frankel, AE, 111<br />

Fredriksson, L, 010, 015<br />

Fu, Q, 038<br />

Furlan, F, 068<br />

Gao, S, 071, 104<br />

Gao, Y, 010<br />

Gårdsvoll, H, 002, 034, 047, 054,<br />

107, 108, 109<br />

Gavard, J, 033<br />

Gerasi, L, 035<br />

Geyer, M, 010<br />

Gils, A, 112, 113<br />

Godiksen, S, 078<br />

González, B, 040<br />

Green, KA, 032<br />

Grosser, M, 098<br />

Gueler, F, 090<br />

Guenther, A, 039<br />

Guo, Y, 036, 093<br />

Gutkind, JS, 033<br />

Hägglöf, P, 023, 091<br />

Hagström, E, 036<br />

Hajjar, KA, 009<br />

Haller, H, 085, 090<br />

Hansen, HD, 115<br />

Hansen, LV, 097, 107<br />

Hansson, G, 070<br />

Harslund, J, 084<br />

Heier, PC, 063<br />

Helgel<strong>and</strong>, L, 095<br />

Henderikx, P, 045<br />

Henic, E, 070<br />

Herry, C, 020<br />

Higgins, PJ, 016<br />

Hill, M, 013, 042<br />

Hillig, T, 034<br />

Holmbeck, K, 033<br />

Høyer-Hansen, G, 034, 047, 070,<br />

100, 109, 116<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 129


Huang, M, 048<br />

Huber, K, 094<br />

Hultman, K, 067<br />

Iaccarino, I, 087<br />

Ibba-Manneschi, L, 065<br />

Ihara, H, 007<br />

Illemann, M, 043, 095, 101<br />

Ingvarsen, S, 034<br />

Jackson, C, 072<br />

Jacobsen, B, 108<br />

Jardí, M, 040, 089<br />

Jensen, JK, 004, 114, 115<br />

Jensen, L, 060<br />

Jerczynska, H, 076<br />

Jern, C, 067<br />

Jögi, A, 047, 109, 116<br />

Johnsen, M, 105<br />

Jørgensen, TJD, 107<br />

Juncker-Jensen, A, 096<br />

Kamikubo, Y, 069<br />

Kanse, SM, 039<br />

Karlsson-Li, S, 023<br />

Kaun, C, 094<br />

Kietzmann, T, 011<br />

Kim, Y, 013, 042<br />

Kirsch, T, 090<br />

Kiyan, J, 085<br />

Kjelgaard, S, 049<br />

Kjems, J, 050<br />

Kjøller, L, 034<br />

Kojima, S, 064<br />

Koschelnick, Y, 031<br />

Kossiak<strong>of</strong>f, A, 103<br />

Kotzsch, M, 098<br />

Kreek, MJ, 022<br />

Kristjansen, PEG, 046<br />

Krogdahl, A, 071, 104<br />

Krog-Mikkelsen, I, 060<br />

Kruith<strong>of</strong>, EKO, 012<br />

Kugler, MC, 013<br />

Kui, L, 081<br />

Lademann, U, 027<br />

Ladner, RC, 045<br />

Laerum, OD, 095, 101, 105<br />

Larsen, L, 027<br />

Larsson, G, 029<br />

Lawrence, DA, 010, 056<br />

Lee, JA, 080<br />

Lengyel, E, 103<br />

Leppla, SH, 046, 111<br />

Leung, HJ, 111<br />

Ley, A, 045<br />

Li, H, 015<br />

Li, J, 036, 088<br />

Li, J, 017<br />

Li, S, 056<br />

Li, X, 001<br />

Liguori, E, 062<br />

Lillis, AP, 008<br />

Ling, Q, 009<br />

List, K, 024, 025, 079<br />

Liu, S, 046, 111<br />

Llorens, A, 082<br />

Lobov, S, 058, 059<br />

Lochner, JE, 021<br />

Lomas, DA, 023, 091<br />

Longanesi-Cattani, I, 062<br />

López-Alemany, R, 082<br />

Lu, W, 001<br />

Lund, IK, 047, 109, 116<br />

Lund, LR, 032, 043, 047,<br />

101,105,109,116<br />

Luque, T, 082<br />

Luther, T, 098<br />

Lüthi, A, 020<br />

Lyden, D, 009<br />

Mackie, I, 045<br />

Madsen, CD, 003, 051, 066<br />

Madsen, JB, 050<br />

Madsen, DH, 034<br />

Magdolen, V, 037, 098<br />

Maiya, R, 022<br />

Majeed, A, 101<br />

Malisauskas, M, 077<br />

Mallya, M, 091<br />

Malmendal, A, 115<br />

Manch<strong>and</strong>a, N, 055<br />

Mancini, A, 074<br />

Manetti, M, 065<br />

Mann, K, 010<br />

Margheri, F, 065<br />

Markart, P, 039<br />

Mathiasen, L, 114<br />

Maurer, G, 094<br />

Mazzieri, R, 035<br />

Medcalf, R, 067<br />

Meins, M, 020<br />

Meissenheimer, LM, 112, 113<br />

Mengel, M, 090<br />

Meye, A, 098<br />

Meyer, CA, 086<br />

Michos, O, 018<br />

Migliorini, M, 008<br />

Mihaly, J, 031, 044, 063<br />

Mikhailenko, I, 008<br />

Minor, KH, 004, 054<br />

Moayeri, M, 111<br />

Mogami, H, 007<br />

Molinolo, A, 024, 025, 079, 111<br />

Molloy, M, 105<br />

Monard, D, 018, 020, 099<br />

Montag, A, 103<br />

Montuori, N, 026<br />

Moreno, E, 018, 020<br />

Morin, SJ, 086<br />

Morozova-Roche, L, 077<br />

Mortensen, KK, 049<br />

Morty, RE, 039<br />

Muñoz-Cánoves, P, 040, 089<br />

Naa, L, 045<br />

Navarro, P, 073<br />

Neels, JG, 069<br />

Netti, PA, 041<br />

Netzel-Arnett, S, 079<br />

Neville, DM, 046<br />

Nielsen, BS, 043, 095, 101<br />

Nielsen, NC, 115<br />

Nielsen, OL, 084<br />

Nieves-Li, EC, 055<br />

Nilsson, M, 067<br />

Norris, EH, 022<br />

Noskova, V, 070<br />

Ny, T, 036, 077, 081, 088, 091, 093<br />

Nyl<strong>and</strong>er, A, 081<br />

Offenberg, H, 020, 084<br />

Olausson, B, 029, 083<br />

Ol<strong>of</strong>sson, A, 077<br />

Orolicki, S, 018<br />

Ortiz-Zapater, E, 073<br />

Ossowski, L, 041<br />

Østergaard, S, 108<br />

Øvrebø, K, 095<br />

Owen, K, 030<br />

Pabba, M, 077<br />

Pappot, H, 097, 100<br />

Park, J-K, 090<br />

Pass, J, 109<br />

Paul, J, 019<br />

Pawlowska, Z, 076<br />

Pedersen, EDK, 078<br />

Peiró, S, 073<br />

Perron, MJ, 054<br />

Peterson, CB, 004, 054<br />

Pfaffenberger, S, 094<br />

Phillips, RL, 091<br />

Pietras, K, 010<br />

Pirazzoli, V, 051<br />

Pizzo, SV, 008<br />

Pliyev, BK, 014<br />

Ploplis, VA, 028,086<br />

Ploug, M, 002, 032, 047, 054, 097,<br />

107, 108<br />

Polavarapu, R, 006<br />

Potempa, J, 037<br />

Prager, GW, 031, 044, 063<br />

130 X I t h I n t e r n a t i o n a l W o r k s h o p o n


Preissner, KT, 039<br />

Priglinger, U, 063<br />

Prorok, M, 038<br />

Przygodzka, P, 029, 083<br />

Przygodzki, T, 077, 091<br />

Pucci, M, 065<br />

Qiu, D, 030<br />

Quigley, JP, 049<br />

Rabbani, SA, 045<br />

Raben, A, 060<br />

Radjabi, R, 103<br />

Rafii, S, 009<br />

Ragno, P, 026<br />

Rank, F, 043<br />

Ranson, M, 038, 058, 059, 080, 102<br />

Rasch, MG, 109<br />

Real, FX, 073<br />

Rega, G, 094<br />

Ricci, P, 026<br />

Robillard, L, 013<br />

Robinson, S, 008<br />

Roda, O, 073<br />

Rømer, J, 032, 046, 047, 096, 105, 116<br />

Rømer, MU, 027<br />

Rong, S, 090<br />

Rønø, B, 046, 109<br />

Rosales, L, 092<br />

Rossi, G, 026<br />

Rotoli, B, 026<br />

Ruan, J, 009<br />

Ruiz, BH, 092<br />

Ruskowski, M, 045<br />

Rychli, K, 094<br />

Saldanha, R, 105<br />

Salvi, A, 110<br />

Samarakoon, R, 016<br />

S<strong>and</strong>erson-Smith, ML, 038<br />

Sarra Ferraris, GM, 003, 066<br />

Saunders, D, 102<br />

Sawada, K, 103<br />

Scalettar, BA, 021<br />

Schar, CR, 004, 054<br />

Scharschmidt, T, 025<br />

Schmitt, M, 037, 098<br />

Schuttner, LC, 021<br />

Schwartz, BS, 056<br />

Segre, J, 025<br />

Selleri, C, 026<br />

Selzer-Plon, J, 078<br />

Serrano, AL, 089<br />

Serrati, S, 065<br />

Sharp, LK, 023<br />

Shireman, J, 025<br />

Shore, JD, 054<br />

Shushakova, N, 085, 090<br />

Sidenius, N, 003, 051, 052, 066<br />

Sidman, RL, 017<br />

Sieuwerts, A, 098<br />

Skarstein, A, 095<br />

Skov, BG, 097<br />

Sloth, B, 060<br />

Smid, M, 098<br />

Smith, RC, 072<br />

Smolarczyk, K, 075<br />

Snyder, EY, 017<br />

Sommerh<strong>of</strong>f, CP, 037<br />

Song, E, 072<br />

Sørensen, JA, 071, 104<br />

Spangler, E, 021<br />

Spina, R, 062<br />

Stochl, M, 045<br />

Stoppelli, MP, 041, 062, 074, 087<br />

Strickl<strong>and</strong>, DK, 008, 010<br />

Strickl<strong>and</strong>, S, 019, 022<br />

Su, EJ, 010<br />

Suelves, M, 089<br />

Sui, G-Z, 009<br />

Sulniute, R, 088<br />

Suzuki, Y, 007<br />

Szabo, R, 024, 025, 079<br />

Taieb, S, 018<br />

Tang, CH, 013, 042<br />

Té, H, 018<br />

Tengel, T, 029<br />

Tholstrup, T, 060<br />

Tjärnlund-Wolf, A, 067<br />

Tkachuk, N, 085<br />

Tkachuk, S, 085<br />

Tkachuk, VA, 014<br />

Toews, M, 045<br />

Urano, T, 007<br />

Vaillant, C, 018<br />

Vallecillo, AJ, 092<br />

van Gool, R, 045<br />

Vidal, B, 040, 089<br />

Visconte, V, 026<br />

Vocca, I, 041<br />

Vogel, LK, 078<br />

Votta, G, 041, 074<br />

Wagenaar-Miller, RA, 033, 034<br />

Wahlberg, P, 081<br />

Waisman, DM, 057<br />

Walker, MJ, 038<br />

Wang, H, 111<br />

Wei, Y, 013, 042<br />

Wei, Z, 005<br />

Wilczynska, M, 029, 077, 083<br />

Wind, T, 039, 048, 114, 115<br />

Wojciechowski, P, 037<br />

Wojta, J, 094<br />

Wygrecka, M, 039<br />

Xolalpa, W, 092<br />

Xu, N, 105<br />

Xue, A, 072<br />

Xue, M, 072<br />

Yamada, S, 033<br />

Yan, L, 009<br />

Yepes, M, 006, 010<br />

Yuan, C, 048<br />

Yuan, W, 001<br />

Zeller, R, 018<br />

Zhao, G, 048<br />

Zhou, A, 005<br />

Zhou, Y, 022<br />

Zou, G, 001<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 131


Attendees<br />

Ahlskog, Nina<br />

Umeå University<br />

nina.ahlskog@medchem.umu.se<br />

Alfano, Daniela<br />

Institute <strong>of</strong> Genetics & Biophysics<br />

alfano@igb.cnr.it<br />

Almholt, Kasper<br />

Finsen Laboratory<br />

kasper@finsenlab.dk<br />

Alpízar Alpízar, Warner<br />

The Gades Institute - Haukel<strong>and</strong> University<br />

Hospital<br />

awarnercr@yahoo.com<br />

Andersen, Lisbeth Moreau<br />

University <strong>of</strong> Aarhus<br />

lma@mb.au.dk<br />

Andolfo, Annapaola<br />

Fondazione Ifom<br />

annapaola.<strong>and</strong>olfo@ifom-ieo-campus.it<br />

Andreasen, Peter A.<br />

University <strong>of</strong> Aarhus<br />

pa@mb.au.dk<br />

Antalis, Toni<br />

University <strong>of</strong> Maryl<strong>and</strong> School <strong>of</strong> Medicine<br />

tantalis@som.umaryl<strong>and</strong>.edu<br />

Ardite, Esther<br />

Center for Genomic Regulation (CRG)<br />

esther.ardite@crg.es<br />

Baker, Mark<br />

Australian Proteome Analysis Facility<br />

mbaker@proteome.org.au<br />

Balsara, Rashna<br />

University <strong>of</strong> Notre Dame<br />

rbalsara@nd.edu<br />

Beaufort, Nathalie<br />

KliFo der Frauenklinik, TU München<br />

nbeaufortgbb@yahoo.fr<br />

Bednarek, Radoslaw<br />

Polish Academy <strong>of</strong> Science - Lodz<br />

rbednarek@cbm.pan.pl<br />

Behrendt, Niels<br />

Finsen Laboratory, Rigshospitalet<br />

niels.behrendt@finsenlab.dk<br />

Binder, Bernd R.<br />

Medical University <strong>of</strong> Vienna<br />

bernd.binder@univie.ac.at<br />

Blasi, Francesco<br />

Fondazione Ifom<br />

francesco.blasi@ifom-ieo-campus.it<br />

Bødker, Julie Støve<br />

University <strong>of</strong> Aarhus<br />

jsb@mb.au.dk<br />

Brunner, Patrick<br />

Medical University <strong>of</strong> Vienna<br />

patrick.brunner@meduniwien.ac.at<br />

Bruno, Katharina<br />

University <strong>of</strong> Chicago<br />

kbruno@uchicago.edu<br />

Bugge, Thomas<br />

National Institutes <strong>of</strong> Health<br />

tbugge@mail.nih.gov<br />

Chapman, Hal<br />

University <strong>of</strong> California San Francisco<br />

hal.chapman@ucsf.edu<br />

Cierniewski, Czeslaw<br />

Polish Academy <strong>of</strong> Sciences<br />

cciern@zdn.am.lodz.pl<br />

Cochran, Blake<br />

University <strong>of</strong> Wollongong<br />

blake@uow.edu.au<br />

Croucher, David<br />

University <strong>of</strong> Wollongong<br />

d.croucher@garvan.org.au<br />

Dano, Keld<br />

Finsen Laboratory<br />

kelddano@dadlnet.dk<br />

Degryse, Bernard<br />

Fondazione Ifom<br />

degryse.bernard@hsr.it<br />

Del Rosso, Mario<br />

University <strong>of</strong> Florence<br />

delrosso@unifi.it<br />

Devy, Laetitia<br />

DYAX SA<br />

ldevy@dyax.com<br />

Diaz Ramos, Mª Angeles<br />

Institut Investigacio BiomedicaBellvitge<br />

madiaz@idibell.org<br />

Dumler, Inna<br />

Hannover Medical School<br />

dumler.inna@mh-hannover.de<br />

Dupont, Daniel<br />

University <strong>of</strong> Aarhus<br />

dmd@mb.au.dk<br />

Eden, Gabriele<br />

Fondazione Ifom<br />

gabriele.eden@ifom-ieo-campus.it<br />

132 X I t h I n t e r n a t i o n a l W o r k s h o p o n


Ehart, Monika<br />

Medical University <strong>of</strong> Vienna<br />

monika.ehart@meduniwien.ac.at<br />

Ehnman, Monika<br />

Ludwig Institute for Cancer Research<br />

monika.ehnman@licr.ki.se<br />

Ellis, Vincent<br />

University <strong>of</strong> East Anglia<br />

v.ellis@uea.ac.uk<br />

Fayard, Berengere<br />

Friedrich Miescher Institute for Biomedical<br />

Research<br />

berengere.fayard@fmi.ch<br />

Fibbi, Gabriella<br />

University <strong>of</strong> Florence<br />

fibbi@unifi.it<br />

Fish, Richard<br />

Geneva University Medical Faculty <strong>and</strong> University<br />

Hospital<br />

Richard.Fish@medecine.unige.ch<br />

Franco, Paola<br />

Institute <strong>of</strong> Genetics & Biophysics<br />

franco@igb.cnr.it<br />

Friberger, Petfer<br />

DiA-Service/American Diagnostica<br />

petter.friberger@dia-service.se<br />

Gårdsvoll, Henrik<br />

Finsen Laboratory, Rigshospitalet<br />

gvoll@finsenlab.dk<br />

Gils, Ann<br />

Laboratory for Pharmaceutical <strong>Biology</strong>, KULeuven<br />

ann.gils@pharm.kuleuven.be<br />

Godiksen, Sine<br />

University <strong>of</strong> Copenhagen<br />

sinego@imbg.ku.dk<br />

Guo, YongZhi<br />

Umeå University<br />

Yong-zhi.guo@medchem.umu.se<br />

Hägglöf, Peter<br />

University <strong>of</strong> Cambridge<br />

pmh43@cam.ac.uk<br />

Hajjar, Katherine<br />

Weill Cornell Medical College<br />

khajjar@med.cornell.edu<br />

Harslund, Jakob<br />

University <strong>of</strong> Copenhagen<br />

jhar@life.ku.dk<br />

Henic, Emir<br />

University Hospital Lund<br />

emir.henic@med.lu.se<br />

Høyer-Hansen, Gunilla<br />

Finsen Laboratory<br />

gunilla@finsenlab.dk<br />

Huang, Mingdong<br />

Chinese Academy <strong>of</strong> Sciences<br />

mhuang@fjirsm.ac.cn<br />

Hultman, Karin<br />

Institution <strong>of</strong> Neurosciences <strong>and</strong> Physiology<br />

karin.hultman@neuro.gu.se<br />

Illemann, Martin<br />

Finsen Laboratory<br />

millemann@finsenlab.dk<br />

Jacobsen, Benedikte<br />

Finsen Laboratory, Rigshospitalet<br />

bjacobsen@finsenlab.dk<br />

Jensen, Lotte<br />

University <strong>of</strong> Copenhagen<br />

lje@life.ku.dk<br />

Jerczynska, Hanna<br />

Medical University <strong>of</strong> Lodz - Pol<strong>and</strong><br />

hanuka@zdn.am.lodz.pl<br />

Jerke, Uwe<br />

Franz-Volhard-Klink, Charite-Buch Berlin<br />

uwe.jerke@charite.de<br />

Jern, Christina<br />

Institution <strong>of</strong> Neurosciences <strong>and</strong> Physiology<br />

christina.jern@neuro.gu.se<br />

Juncker-Jensen, Anna<br />

Finsen Laboratory<br />

ajjensen@finsenlab.dk<br />

Kietzmann, Thomas<br />

University <strong>of</strong> Kaiserslautern<br />

tkietzm@gwdg.de<br />

Kinnby, Bertil<br />

Malmö University College<br />

Bertil.Kinnby@od.mah.se<br />

Kojima, Soichi<br />

<strong>Molecular</strong> <strong>Cellular</strong> Pathology Research Unit,<br />

RIKEN<br />

skojima@postman.riken.go.jp<br />

Kruith<strong>of</strong>, Egbert<br />

University Hospital <strong>of</strong> Geneva<br />

egbert.kruith<strong>of</strong>@hcuge.ch<br />

Kusch, Angelika<br />

Charite-Campus Buch, Dept <strong>of</strong> <strong>Molecular</strong> <strong>and</strong><br />

Clinical Cardiology<br />

angelika.kusch@charite.de<br />

Lademann, Ulrik<br />

University <strong>of</strong> Copenhagen<br />

ul@life.ku.dk<br />

Laerum, Ole Didrik<br />

The Gade Institute, University <strong>of</strong> Bergen<br />

ole.laerum@gades.uib.no<br />

Lawrence, Daniel<br />

University <strong>of</strong> Michigan Medical School<br />

dlawrenc@med.umich.edu<br />

Lee, Jodi<br />

University <strong>of</strong> Wollongong<br />

jal31@uow.edu.au<br />

Li, Jianxue<br />

BIDMC, Harvard Medical School<br />

jli7@caregroup.harvard.edu<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 133


Li, Jinan<br />

Umeå University<br />

jinan.li@medchem.umu.se<br />

Li, Shih-Hon<br />

University <strong>of</strong> Illinois at Urbana-Champaign<br />

sli3@uiuc.edu<br />

Lillis, Anna<br />

University <strong>of</strong> Maryl<strong>and</strong> School <strong>of</strong> Medicine<br />

alillis@som.umaryl<strong>and</strong>.edu<br />

Liu, Kui<br />

Umeå University<br />

kui.liu@medchem.umu.se<br />

Lobov, Sergei<br />

University <strong>of</strong> Wollongong<br />

sergei@uow.edu.au<br />

Lochner, Janis<br />

Lewis & Clark College<br />

lochner@lclark.edu<br />

Longanesi Cattani, Immacolata<br />

National Cancer Institute <strong>of</strong> Naples - Italy<br />

immalonganesi@libero.it<br />

Lu, Wuyuan<br />

University <strong>of</strong> Maryl<strong>and</strong><br />

luw@umbi.umd.edu<br />

Lund, Ida K.<br />

Finsen Laboratory<br />

ikl@finsenlab.dk<br />

Lund, Leif R.<br />

Finsen Laboratory<br />

lund@inet.uni2.dk<br />

Madsen, Chris<br />

Fondazione Ifom<br />

chris.madsen@ifom-ieo-campus.it<br />

Magdolen, Viktor<br />

Clinical Research Unit, TU München<br />

viktor-magdolen@lrz.tum.de<br />

Maiya, Rajani<br />

The Rockefeller University<br />

rmaiya@mail.rockefeller.edu<br />

Manch<strong>and</strong>a, Naveen<br />

University <strong>of</strong> Illinois at Urbana-Champaign<br />

manch<strong>and</strong>@uiuc.edu<br />

Markart, Philipp<br />

University <strong>of</strong> Giessen Lung Center<br />

philipp.markart@innere.med.uni-giessen.de<br />

Meins, Marita<br />

Friedrich Miescher Institute for Biomedical<br />

Research<br />

mmeins@fmi.ch<br />

Meissenheimer, Lester<br />

KULeuven: Lab Pharmaceutical <strong>Biology</strong><br />

lester.meissenheimer@pharm.kuleuven.be<br />

Mihaly, Judit<br />

Medical University <strong>of</strong> Vienna<br />

judit.mihaly@meduniwien.ac.at<br />

Monard, Denis<br />

Friedrich Miescher Institute for Biomedical<br />

Research<br />

monard@fmi.ch<br />

Navarro, Pilar<br />

Institut Municipal dInvestigació Mèdica<br />

pnavarro@imim.es<br />

Nielsen, Boye Schnack<br />

Finsen Laboratory, Rigshospitalet<br />

schnack@finsenlab.dk<br />

Nieves-Li, Evelyn<br />

University <strong>of</strong> Illinois at Urbana-Champaign<br />

enieves@uiuc.edu<br />

Nuttall, Robert<br />

Dalhousie University<br />

r.nuttall@dal.ca<br />

Ny, Tor<br />

Umeå University<br />

tor.ny@medchem.umu.se<br />

Olausson, Björn<br />

Umeå University<br />

bjorn.olausson@medchem.umu.se<br />

Pabba, Mohan<br />

Umeå University<br />

pabba.mohan@medchem.umu.se<br />

Pappot, Helle<br />

Rigshospitalet 5073<br />

pappot@rh.regionh.dk<br />

Paul, Justin<br />

The Rockefeller University<br />

jpaul@rockefeller.edu<br />

Peterson, Cynthia<br />

University <strong>of</strong> Tennessee<br />

cbpeters@utk.edu<br />

Pirazzoli, Valentina<br />

Fondazione Ifom<br />

valentina.pirazzoli@ifom-ieo-campus.it<br />

Pliyev, Boris<br />

Moscow State University<br />

bpliyev@cardio.ru<br />

Ploug, Michael<br />

Finsen Laboratory<br />

m-ploug@finsenlab.dk<br />

Prager, Gerald<br />

Medical University <strong>of</strong> Vienna<br />

gerald.prager@meduniwien.ac.at<br />

Przygodzka, Patrycja<br />

Umeå University<br />

patrycja.przygodzka@medchem.umu.se<br />

Przygodzki, Tomasz<br />

Umeå University<br />

tomasz.przygodzki@medchem.umu.se<br />

Ragno, Pia<br />

University <strong>of</strong> Salerno<br />

pragno@unisa.it<br />

134 X I t h I n t e r n a t i o n a l W o r k s h o p o n


Ranson, Marie<br />

University <strong>of</strong> Wollongong<br />

mranson@uow.edu.au<br />

Rasch, Morten<br />

Finsen Laboratory<br />

mrasch@finsenlab.dk<br />

Resnati, Massimo<br />

S.Raffaele Hospital- DIBIT<br />

resnati.massimo@hsr.it<br />

Rønø, Birgitte<br />

Finsen Laboratory<br />

brono@finsenlab.dk<br />

Salvi, Aless<strong>and</strong>ro<br />

University <strong>of</strong> Brescia<br />

asalvi@med.unibs.it<br />

Sarra Ferraris, Gian Maria<br />

Fondazione Ifom<br />

gianmaria.sarraferraris@ifom-ieo-campus.it<br />

Schmitt, Manfred<br />

Clinical Research Unit, Dept. Obstetrics <strong>and</strong><br />

Gynecology<br />

manfred.schmitt@LRZ.tum.de<br />

Schwartz, Brad<br />

University <strong>of</strong> Illinois at Urbana-Champaign<br />

comuc@med.uiuc.edu<br />

Sidenius, Nicolai<br />

Fondazione Ifom<br />

nicolai.sidenius@ifom-ieo-campus.it<br />

Stoppelli, Maria Patrizia<br />

Institute <strong>of</strong> Genetics & Biophysics<br />

stoppell@igb.cnr.it<br />

Strickl<strong>and</strong>, Dudley<br />

University <strong>of</strong> Maryl<strong>and</strong> School <strong>of</strong> Medicine<br />

dstrickl<strong>and</strong>@som.umaryl<strong>and</strong>.edu<br />

Sulniute, Rima<br />

Umeå University<br />

rima.sulniute@medchem.umu.se<br />

Suzuki, Yuko<br />

Hamamatsu University School <strong>of</strong> Medicine<br />

seigan@hama-med.ac.jp<br />

Szabo, Roman<br />

National Institutes <strong>of</strong> Health<br />

rszabo@nidcr.nih.gov<br />

Tkachuk, Sergey<br />

Hannover Medical School<br />

Tkatchouk.Sergei@mh-hannover.de<br />

Urano, Tetsumei<br />

Hamamatsu University School <strong>of</strong> Medicine<br />

uranot@hama-med.ac.jp<br />

Vidal, Berta<br />

Center for Genomic Regulation (CRG)<br />

berta.vidal@crg.es<br />

Vogel, Lotte Katrine<br />

University <strong>of</strong> Copenhagen<br />

vogel@imbg.ku.dk<br />

Wahlberg, Patrik<br />

Umeå University<br />

patrik.wahlberg@medchem.umu.se<br />

Wei, Ying<br />

University <strong>of</strong> California San Francisco<br />

ying.wei@ucsf.edu<br />

Wikström, Clas<br />

Umeå University<br />

clas.wikstrom@medchem.umu.se<br />

Wilczynska, Malgorzata<br />

Umeå University<br />

Malgorzata.Wilczynska@medchem.umu.se<br />

Wojta, Johann<br />

Medical University <strong>of</strong> Vienna<br />

johann.wojta@meduniwien.ac.at<br />

Wygrecka, Malgorzata<br />

University <strong>of</strong> Giessen Lung Center<br />

malgorzata.wygrecka@innere.med.uni-giessen.de<br />

Xolalpa, Wendy<br />

Instituto de Investigaciones Biomedicas UNAM<br />

wendyxolalpa@yahoo.com.mx<br />

Xue, Aiqun<br />

University <strong>of</strong> Sydney Royal North Shore Hospital<br />

aiqunn@med.usyd.edu.au<br />

Yepes, Manuel<br />

Emory University<br />

myepes@emory.edu<br />

Zhou, Aiwu<br />

University <strong>of</strong> Cambridge<br />

awz20@cam.ac.uk<br />

<strong>Molecular</strong> & <strong>Cellular</strong> <strong>Biology</strong> <strong>of</strong> <strong>Plasminogen</strong> <strong>Activation</strong> 135


Notes<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

136 X I t h I n t e r n a t i o n a l W o r k s h o p o n

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!