31.10.2013 Views

PhD Thesis - Biologisk Institut

PhD Thesis - Biologisk Institut

PhD Thesis - Biologisk Institut

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

F A C U L T Y O F S C I E N C E<br />

U N I V E R S I T Y O F C O P E N H A G E N<br />

<strong>PhD</strong> <strong>Thesis</strong><br />

Sine Godiksen<br />

Towards an Understanding of the Role of<br />

Matriptase in Normal Physiology<br />

Academic advisors: Lone Rønnov-Jessen and Lotte K. Vogel<br />

Submitted: 17/09/2012


Towards an Understanding of the Role of<br />

Matriptase in Normal Physiology<br />

Submitted to:<br />

The Graduate School of Science, Faculty of Science<br />

Department of Biology, University of Copenhagen, Denmark<br />

For the <strong>PhD</strong> Degree<br />

by<br />

Sine Godiksen<br />

Department of Biology, University of Copenhagen<br />

Universitetsparken 13, 2100 København Ø<br />

and<br />

Department of Cellular and Molecular Medicine, University of Copenhagen<br />

Blegdamsvej 3, 2200 København N<br />

September 2012


Preface<br />

This thesis is submitted to the Faculty of Science, University of Copenhagen, Denmark as basis for<br />

obtaining the <strong>PhD</strong> degree. I received a personal scholarship from the Faculty of Science in October<br />

2007 and started my <strong>PhD</strong> studies in January 2008. My <strong>PhD</strong> thesis entitled: ”Towards an<br />

understanding of the role of matriptase in normal physiology” is based on the work performed<br />

from November 2009 to August 2011 in the Lab of Lotte Vogel, Department of Cellular and<br />

Molecular Medicine, Faculty of Health and Medical Sciences, University of Copenhagen, with Dr.<br />

scient Lone Rønnov-Jessen, Department of Biology, Faculty of Science, University of Copenhagen<br />

as faculty supervisor and Lotte Vogel as supervisor.<br />

Additionally, I have included work performed in the Lab of <strong>PhD</strong> Thomas H. Bugge, National<br />

<strong>Institut</strong>es of Health, Bethesda, MD, USA under a leave of absence from my <strong>PhD</strong> studies from<br />

September 2011 – April 2012 (paper II).<br />

The data obtained for this thesis have been presented both at national and international scientific<br />

meetings and the results have resulted in three scientific papers; two published manuscripts and a<br />

manuscript in progress (see list of papers).<br />

My research was financially supported by the Faculty of Science, Danish Cancer Society, the<br />

Augustinus Foundation, Købmand Kristian Kjær og Hustrus Foundation - the Kjær-Foundation,<br />

Dagmar Marshall´s Foundation, Snedkermester Sophus Jacobsen og Hustru Astrid Jacobsen´s<br />

Foundation, Grosserer Valdemar Foersom og Hustru Thyra Foersom´s Foundation, Fabrikant Einar<br />

Willumsens Mindelegat.<br />

Sine Godiksen<br />

Cand. Scient.<br />

Copenhagen, September 2012


Table of Contents<br />

ACKNOWLEDGEMENTS .......................................................................................................... 3<br />

LIST OF PAPERS ...................................................................................................................... 4<br />

ABSTRACT ............................................................................................................................... 5<br />

DANSK RESUMÉ ...................................................................................................................... 7<br />

LIST OF ABBREVIATIONS ....................................................................................................... 9<br />

INTRODUCTION ...................................................................................................................... 11<br />

AIMS AND OBJECTIVES ........................................................................................................ 13<br />

BACKGROUND ....................................................................................................................... 14<br />

Membrane-anchored serine proteases ................................................................................................................14<br />

Matriptase ...........................................................................................................................................................16<br />

Matriptase expression and structure .......................................................................................................................... 16<br />

Regulation of matriptase ......................................................................................................................................18<br />

Matriptase inhibitors .................................................................................................................................................. 18<br />

Matriptase proteolytic processing and activation ...................................................................................................... 20<br />

Prostasin ..............................................................................................................................................................22<br />

Physiological functions of matriptase and its role in pathological processes ........................................................23<br />

Matriptase is crucial for epidermal integrity .............................................................................................................. 23<br />

The matriptase-prostasin proteolytic pathway(s) in the epidermis ........................................................................... 24<br />

Matriptase has an important role in epithelial homeostasis ...................................................................................... 25<br />

Importance of matriptase regulation in vivo .............................................................................................................. 26<br />

Matriptase in carcinogenesis ...................................................................................................................................... 27<br />

TECHNICAL CONSIDERATIONS............................................................................................ 29<br />

Cell system ...........................................................................................................................................................29<br />

Protease pull down assays ...................................................................................................................................30<br />

Biotinylation assays..............................................................................................................................................32<br />

RESULTS ................................................................................................................................ 35


Paper I ..................................................................................................................................................................36<br />

Paper II .................................................................................................................................................................47<br />

Paper III ................................................................................................................................................................65<br />

DISCUSSION AND PERSPECTIVES ...................................................................................... 85<br />

REFERENCES ......................................................................................................................... 91<br />

SUPPLEMENTARY I ............................................................................................................. 105<br />

2


Acknowledgements<br />

During the process of gaining experimental data and knowledge to complete this thesis and<br />

obtain my <strong>PhD</strong> there are many people whom I would like to thank. Without their continuous help,<br />

encouragement and support, this work would not have been possible. I would hereby like to<br />

express my gratitude to:<br />

Lone Rønnov-Jessen for taking on the job as my supervisor in difficult times. Thank you for your<br />

great support and for your invaluable guidance in my writing process.<br />

Lotte K. Vogel for giving me the opportunity to finish my <strong>PhD</strong> studies, and for your support<br />

throughout my <strong>PhD</strong>.<br />

Thomas Bugge for giving me a very interesting and inspiring stay in his lab in Bethesda and for his<br />

excellent scientific guidance.<br />

Jan K. Jensen for valuable discussions throughout my <strong>PhD</strong> studies.<br />

All of my colleagues at ICMM, University of Copenhagen and NIDCR, National <strong>Institut</strong>es of Health<br />

(NIH), for a very pleasant, helpful and inspiring environment.<br />

Stine in particular, for support, scientific discussions, and for establishing the getaway to NIH.<br />

Karen Skriver for your interest and deeply valued support throughout my <strong>PhD</strong>.<br />

Annette Storgaard for your professionalism and for your guidance.<br />

Family and friends for your constant support and encouragements.<br />

And finally my beloved Peter, “tak for mad” and for being the most supportive and understanding<br />

“kæreste” anyone could ask for.<br />

3


List of papers<br />

Papers included in the thesis:<br />

Paper I<br />

Transport via the transcytotic pathway makes prostasin available as substrate for matriptase.<br />

Stine Friis, Sine Godiksen, Jette Bornholdt, Joanna Selzer-Plon, Hanne Borger Rasmussen, Thomas<br />

H. Bugge, Chen‐Yong Lin and Lotte K. Vogel<br />

Paper II<br />

Reduced prostasin (CAP1/PRSS8) activity eliminates HAI-1 and HAI-2 deficiency-associated<br />

developmental defects by preventing matriptase activation. Roman Szabo, Katiuchia Uzzun<br />

Sales, Peter Kosa, Natalia A. Shylo, Sine Godiksen, Karina K. Hansen, Stine Friis, Silvio Gutkind,<br />

Lotte K. Vogel, Edith Hummler, Eric Camerer, and Thomas H. Bugge<br />

Paper III<br />

Novel assay for detection of active matriptase.<br />

Sine Godiksen and Lotte Vogel. Other authors to be included.<br />

Papers not included in the thesis:<br />

Bornholdt J, Friis S, Godiksen S, Poulsen SS, Santoni‐Rugiu E, Bisgaard HC, Bowitz Lothe IM, Ikdahl<br />

T, Tveit KM, Johnson E, Kure E, Vogel LK. The level of claudin‐7 is reduced as an early event in<br />

colorectal carcinogenesis. BMC Cancer, 2011, Feb 10; 11:65.<br />

4


Abstract<br />

Matriptase is a type II membrane-anchored serine protease essential for epithelial integrity and<br />

with implication in carcinogenesis. Matriptase is co-expressed in epithelial cells of most tissues<br />

with the protease prostasin and their mutual inhibitor hepatocyte growth factor activator<br />

inhibitor 1 (HAI-1). All three proteins have crucial functions for epidermal integrity; and in the<br />

epidermis matriptase acts upstream of prostasin and is required for its activation.<br />

The difference between the subcellular locations of matriptase and prostasin has so far been an<br />

enigma for matriptase-prostasin interaction. By mapping the subcellular trafficking of matriptase,<br />

prostasin and HAI-1, we demonstrate that the basolateral plasma membrane is part of the<br />

subcellular itinerary of both proteases and thus constitutes a location for interaction. In Caco-2<br />

cells, zymogen matriptase is routed to the basolateral plasma membrane where it becomes<br />

activation site cleaved and activated. At steady state, prostasin is found to locate mainly to the<br />

apical plasma membrane although a minor fraction of prostasin can be detected at the<br />

basolateral plasma membrane. We show that prostasin is present in its active form on both the<br />

apical and basolateral plasma membrane and that prostasin as well as HAI-1 are endocytosed<br />

from the basolateral plasma membrane and transcytosed to the apical plasma membrane.<br />

Activation of matriptase on the basolateral plasma membrane is followed by efficient<br />

internalization as a matriptase-HAI-1 complex.<br />

Deregulated or unopposed matriptase activity causes a perturbation of several biological<br />

functions e.g. developmental defects in mice. However, the molecular mechanisms behind these<br />

effects are poorly understood. We performed a genetic epistatic analysis to identify new<br />

components of matriptase-dependent pathways in embryogenesis. We show that prostasin is an<br />

indispensable component of the matriptase-dependent proteolytic cascade that causes early<br />

embryonic lethality in mice. In placental tissue prostasin acts upstream of matriptase and is<br />

required for the activation of matriptase. During embryonic development both HAI-1 and HAI-2<br />

are essential inhibitors of matriptase. In this study we moreover find that both HAI-1 and HAI-2<br />

are indispensable inhibitors of prostasin in placental tissue.<br />

Matriptase activity is a tightly regulated protease; zymogen and HAI-1-complexed matriptase is<br />

abundantly present in most epithelial cell lines and tissues. There are no specific substrates or<br />

inhibitors of matriptase, which has led to challenges in detecting active matriptase. In order to<br />

assess the location and relative amount of active matriptase as compared to the total amount, we<br />

have established an assay for detection of active matriptase. This assay is based on a<br />

chloromethyl ketone peptide inhibitor with a predicted substrate sequence of matriptase. We<br />

show that active matriptase is present on the basolateral plasma membrane of Caco-2 cells and<br />

merely comprise a fraction of total matriptase. The conversion of zymogen matriptase to active<br />

matriptase can be induced by exposure to slight acidic conditions as well as the levels of HAI-1-<br />

complexed matriptase and active matriptase rises under these conditions in Caco-2 cells. Labeling<br />

of Caco-2 cells with the chloromethyl ketone peptide inhibitor delayed matriptase-HAI-1 complex<br />

formation. In addition to the peptide inhibitor being unable to bind HAI-1-complexed matriptase,<br />

we show that the matriptase zymogen has an intrinsic activity that enables it to bind the<br />

chloromethyl ketone peptide inhibitor. Moreover, this assay can be easily modified to detect<br />

5


active matriptase in other cell system, and we show that cultured primary murine keratinocytes<br />

contain low levels of peptidiolytic active matriptase.<br />

Taken together, these results have improved our knowledge of the interplay between matriptase<br />

and prostasin towards a understanding of these components roles in epithelial integrity and<br />

present an assay for detection of active matriptase.<br />

6


Dansk resumé<br />

Matriptase er en type II membranbunden serinprotease, der er essentiel for integriteten af<br />

epitelvæv og er endvidere involveret i udvikling af kræft. Matriptase ekspreseres sammen med<br />

proteasen prostasin og deres fælles proteasehæmmer hepatocyte growth factor activator<br />

inhibitor 1 (HAI-1) af epitelceller i mange væv. Alle tre proteiner har en vital funktion for<br />

integriteten af epidermis, og matriptase er nødvendig for aktivering af prostasin i dette væv.<br />

Forskelle i subcellulær lokalisering mellem matriptase og HAI-1 på basolateralmembranen og<br />

prostasin på apikalmembranen af polariserede epitelceller har hidtil været et uafklaret dilemma i<br />

forhold til de to proteasers samspil. Kortlægning af den intracellulære transport af matriptase,<br />

prostasin og HAI-1 i Caco-2 celler angiver den basolaterale plasmamembran som en mulig<br />

subcellulær lokalitet, hvor matriptase og prostasin kan interagere. Zymogen matriptase<br />

exocyteres til basolateralmembranen, hvor proteasen kløves proteolytisk til den aktive form. Ved<br />

”steady state” befinder prostasin sig primært på apikalmembranen, omend en mindre fraktion<br />

kan detekteres på basolateralmembranen. Vi viser, at aktivt prostasin findes på både apikal -og<br />

basolateralmembranen, og at prostasin såvel som HAI-1 internaliseres fra basolateralmembranen<br />

og transcyteres til apikalmembranen. Aktivering af matriptase på basolateralmembranen<br />

efterfølges af internalisering af matriptase i kompleks med HAI-1.<br />

Dereguleret og ukontrolleret matriptase aktivitet er en tilgrundliggende årsag til uligevægt i flere<br />

vigtige biologiske processer, f.eks. udviklingsmæssige defekter i mus.<br />

Vi udførte genetiske analyser i mus for at identificere nye komponenter i matriptase-afhængige<br />

signalveje. Disse resultater viser, at prostasin er en essentiel komponent i den matriptaseafhængige<br />

proteolytiske kaskade, der resulterer i embryonal dødelighed i mus. Vi viser, at<br />

prostasin ligger opstrøms for matriptase og er nødvendig for aktivering af matriptase i murinet<br />

placental væv. Både HAI-1 og HAI-2 er tidligere vist at være vigtige proteasehæmmere af<br />

matriptase i placenta. I dette studie viser vi, at dette også er gældende for prostasin.<br />

Matriptase er højt reguleret på posttranslationel niveau, hvilket indikeres af, at zymogen<br />

matriptase og matriptase i kompleks med HAI-1 let detekteres i mange cellelinier og epitelvæv.<br />

Der er ikke identificeret en specifik inhibitor eller et specifikt substrate for matriptase, hvilket har<br />

betydet, at detektering af aktivt matriptase har været udfordrende.<br />

Vi udviklede et assay til detektering af aktivt matriptase baseret på en klorometylketon<br />

peptidinhibitor med en peptidsekvens, der afspejler en foretruknen substratesekvens for<br />

matriptase.<br />

Vi viser, at aktivt matriptase findes på basolateralmembranen; denne udgør dog kun en fraktion af<br />

total matriptase. Zymogen aktivering af matriptase kan induceres ved et let fald i pH, hvormed<br />

mængden af aktiv matriptase stiger, ligesom kompleksdannelse med HAI-1 induceres under disse<br />

forhold. Ydermere forsinker mærkning med klorometylketon peptidinhibitoren kompleksdannelse<br />

mellem matriptase og HAI-1; tilsvarende er inhibitoren ikke i stand til at mærke matriptase-HAI-1<br />

komplekset. Vi viser endvidere, at zymogen matriptase er i stand til at binde klorometylketone<br />

peptidinhibitoren. Ydermere kan dette assay let modificeres til andre cellesystemer, og vi viser<br />

med dette assay tilstedeværelsen af aktivt matriptase i kulturer af primære murine keratinocytter.<br />

7


Disse resultater har udvidet vores viden om samspillet mellem matriptase and prostasin mod en<br />

forståelse af disse proteases rolle for integriteten af epitel og præsenterer et assay til detektering<br />

af aktiv matriptase.<br />

8


List of Abbreviations<br />

aa<br />

Amino acid<br />

Arg<br />

Arginine<br />

ARIH<br />

Autosomal recessive ichthyosis with hypotrichosis<br />

Asp<br />

Aspartic acid<br />

cDNA<br />

Copy deoxyribonucleic acid<br />

CAP<br />

Channel activating protease<br />

CDCP1 CUB domain-containing protein 1<br />

CUB<br />

Cls/Clr, urchin embryonic growth factor, and bone morphogenic protein-1<br />

DESC<br />

Differentially expressed in squamous cell carcinoma gene<br />

EGFR<br />

Epidermal growth factor receptor<br />

ENaC<br />

Epithelial Na + channel<br />

ER<br />

Endoplasmic reticulum<br />

Gly<br />

Glycine<br />

GPI<br />

Glycosylphosphatidylinositol<br />

Gpld1<br />

GPI-specific phospholipase D1<br />

HAI-1 Hepatocyte growth factor activator inhibitor 1<br />

HAI-2<br />

Hepatocyte growth factor activator inhibitor2<br />

HAT<br />

Human airway trypsin-like<br />

HGF/SF<br />

Hepatocyte growth factor/scatter factor<br />

HGFA<br />

Hepatocyte growth factor activator<br />

His<br />

Histidine<br />

IGFBP-rP1<br />

Insulin-like growth factor binding protein-related protein1<br />

IP<br />

Immunoprecipitation<br />

KD1<br />

N-terminal Kunitz domain<br />

KD2<br />

C-terminal Kunitz domain<br />

kDa<br />

Kilo dalton<br />

LDLR<br />

Low density lipoprotein receptor class A<br />

Lys<br />

Lysine<br />

MAM<br />

Meprin, A5 antigen, and receptor protein phosphatase μ<br />

MANSC<br />

Motif at the N-terminal containing seven cysteines<br />

MDCK<br />

Mardin Darby canine kidney (cell)<br />

mRNA<br />

Messenger ribonucleic acid<br />

MMP-3 Matrix metalloproteinases 3<br />

MSP-1<br />

Macrophage stimulating protein-1<br />

MSPL<br />

Mosaic serine protease large-form<br />

MT-SP1<br />

Membrane type serine protease 1 (matriptase)<br />

NHS<br />

N-hydroxysulfosuccinimide<br />

PAR-2<br />

Protease-activated receptor-2<br />

PDGF<br />

Platelet-derived growth factor<br />

PN1 Protease nexin 1<br />

9


PRSS14 Protease serine S1 family member 14<br />

PS-SCL<br />

Positional scanning synthetic combinatorial library<br />

S1P<br />

Sphingosine 1-phosphate<br />

SA<br />

Signal anchor<br />

SDS-PAGE<br />

Sodium dodecyl sulphate polyacrylamide gel electrophoresis<br />

SEA<br />

Sea urchin sperm protein, enteropeptidase, and agrin<br />

Ser<br />

Serine<br />

SIMA135<br />

Subtractive immunization M(+)HEp3 associated 135 kDa protein<br />

S-NHS-SS-Biotin Sulfosuccinimidyl-2-[biotinamido]ethyl-1,3-dithiopropionate<br />

SPD<br />

Serine protease domain<br />

ST14<br />

Suppressor of tumorigenicity-14<br />

TADG-15<br />

Tumor-associated differentially expressed gene-15<br />

TEER<br />

Transepithelial electrical resistance<br />

TMPRSS<br />

Transmembrane protease, serine<br />

tPA<br />

Tissue plasminogen activator<br />

TTSP<br />

Type II transmembrane serine protease<br />

uPA<br />

Urokinase-type plasminogen activator<br />

VEGFR-2 Vascular endothelial growth factor receptor 2<br />

10


Introduction<br />

Proteolysis is a critical event in many biological processes from simple protein degradation in<br />

nutrient digestion to specific alterations in protein activity in highly regulated protease cascades,<br />

e.g. the blood coagulation system [1]. Formerly serine proteases represented enzymes that were<br />

either secreted or sequestered in cytoplasmic storage organelles awaiting signal-regulated<br />

release, but analysis of the human genome at the turn of the millennium revealed a new subclass<br />

of serine protease; the membrane-anchored serine proteases. With this family, a whole new<br />

range of important biological processes was shown to be regulated by serine proteases [2;3].<br />

Matriptase is a type II membrane-anchored serine protease and is especially interesting due to its<br />

role in carcinogenesis and due to its importance for development and maintenance of epithelial<br />

integrity [4-8]. However, little is known of matriptase´s role at the molecular level in<br />

carcinogenesis. Of importance is the finding that a modest over-expression of wild-type<br />

matriptase in the epidermis of transgenic mice is sufficient to promote spontaneous squamous<br />

cell carcinoma formation. A simultaneous increase in expression of matriptase´s cognate inhibitor<br />

hepatocyte growth factor activator inhibitor 1 (HAI-1) completely negates the oncogenic<br />

phenotype of matriptase over-expression suggesting that the unopposed matriptase activity is the<br />

underlying cause [6].<br />

Studies of matriptase knock-out mice have revealed critical functions for matriptase in<br />

development and maintenance of multiple epithelial tissues, which is in accordance to its<br />

widespread epithelial distribution in both mice and humans [4;5;9-12]. In humans, mutations in<br />

the ST14 gene encoding matriptase cause congenital ichthyosis [13-15]. Matriptase acts upstream<br />

of the glycosylphosphatidylinositol (GPI)-anchored serine protease prostasin in terminal<br />

epidermal differentiation and evidence suggests that an impairment of this cascade is the<br />

underlying cause of the symptoms observed in these patients [13;15-19]. Moreover, inhibition of<br />

matriptase is an essential function of HAI-1 in maintaining epidermal homeostasis in mice and<br />

zebrafish, and HAI-1 has been shown to inhibit both matriptase and prostasin in cultured human<br />

keratinocytes [20-22].<br />

In contrast to most enzymatic reactions, proteolysis is a “one-way” process, which requires strict<br />

regulation of protease activity. Spatial distribution and confinement of proteases to specific<br />

cellular compartments is one way to ensure that proteolytic processing occurs under the correct<br />

conditions. Matriptase, prostasin and HAI-1 are co-expressed in polarized epithelial cells and HAI-<br />

1 is an inhibitor of both proteases [9;11;21;23;24]. Polarized eepithelial cells have a plasma<br />

membrane that is divided into an apical compartment and a basolateral compartment. Both HAI-1<br />

and matriptase is located at the basolateral plasma membrane in numerous epithelial tissues and<br />

cell types whereas, prostasin is located at the apical plasma membrane [11;23-26]. This difference<br />

in spatial distribution is in conflict with the proposed matriptase-prostasin proteolytic cascade<br />

[16].<br />

We have earlier shown in a recombinant system that HAI-1 is transported to the basolateral<br />

plasma membrane. After endocytosis, a fraction of HAI-1 is transcytosed across the polarized cell<br />

to the apical membrane compartment, suggesting how HAI-1 can access and inhibit both<br />

matriptase and prostasin (supplementary I; [27]). However, it is still uncertain how matriptase is<br />

11


able to activate prostasin in light of their different subcellular localization. Nevertheless, the<br />

global co-expression of matriptase and prostasin could indicate that this proteases cascade has a<br />

general role in maintaining epithelial integrity. Still, it remains unclear what substrates and<br />

pathways are involved in matriptase-dependent epithelial homeostasis, and if prostasin is a global<br />

downstream target for matriptase.<br />

The life cycle of matriptase is very complex and tightly regulated. Matriptase is believed to have a<br />

role as a protease at the pinnacle of protease cascades due to its ability to autoactivate [28;29].<br />

Matriptase activation and inhibition by HAI-1 is shown to be intimately linked and has lead to the<br />

hypothesis that matriptase only exist in its free active form for a limited period and therefore only<br />

has a narrow time window to act on its substrate(s) before being pacified by inhibitor complex<br />

formation [21;30;31]. Thus, experimental tools for detection of active matriptase are highly<br />

desirable.<br />

Hence, matriptase is an important serine protease essential for epithelial integrity and with<br />

implications in cancinogenesis. Still, it remains to be elucidated where in the polarized cell<br />

matriptase is activated and able to act on downstream substrate(s). Moreover, identification of<br />

additional components of matriptase-dependent proteolytic pathways would provide valuable<br />

insights into the biology of matriptase-dependent biological processes.<br />

This thesis is written as a synopsis and gives a brief introduction to membrane-anchored serine<br />

proteases and matriptase in particular. Mechanisms for matriptase activation and inhibition will<br />

be described with an introduction of matriptase´s most important inhibitors, HAI-1 and HAI-2, and<br />

its epidermal downstream target prostasin. Finally, a description of matriptase in relation to its<br />

physiological functions and its role in pathological processes will be given.<br />

The background section is followed by aims and objectives of this thesis and technical<br />

considerations. The results obtained are enclosed as two published papers and one manuscript in<br />

progress. Finally, the results will be discussed in relation to previous studies.<br />

12


Aims and objectives<br />

The overall aim of this study was to gain a more comprehensive understanding of matriptase in<br />

epithelial biology.<br />

We wanted to delineate how matriptase can act as an upstream activator of prostasin in<br />

epidermal terminal differentiation, despite their different subcellular localization in polarized<br />

epithelial cells at steady state, by mapping the intracellular trafficking of matriptase, prostasin and<br />

HAI-1 (paper I). We furthermore wished to determine where in the polarized cell matriptase is<br />

activated and able to cleave its substrates (paper I and paper III). More so, we wanted to establish<br />

an assay for detection of active matriptase on the surface of living cells (paper III). One way of<br />

elucidating the physiological functions of proteins is by identification of interaction partners,<br />

upstream regulators and downstream effectors. Accordingly, we aimed at identifying new<br />

components of matriptase-dependent proteolytic pathways critical for embryogenesis in mice<br />

(paper II).<br />

13


Background<br />

To understand the importance of matriptase and the relationship between matriptase, prostasin<br />

and their inhibitor HAI-1, this chapter gives an introduction to these proteins as well as the<br />

mechanisms of matriptase activation. Afterwards, I will outline important physiological functions<br />

of matriptase, its interplay with its inhibitors and prostasin in these processes, and finally review<br />

the role of matriptase in cancer.<br />

Membrane-anchored serine proteases<br />

Proteases are enzymes that conduct proteolysis and were originally discovered a century ago as<br />

enzymes involved in nutrient digestion. Sequencing of the human genome at the turn of the<br />

millennium revealed a total 569 proteases in humans. On the basis of the mechanism of catalysis,<br />

proteases are classified into five distinct classes: aspartic, metallo, cysteine, serine and threonine<br />

proteases [32].<br />

Serine proteases are one of the largest and most conserved multigene proteolytic families. This<br />

protease family includes the well-known digestive enzymes trypsin and chymotrypsin. Serine<br />

proteases are found in a variety of tissues and body fluids with well characterized roles in diverse<br />

cellular functions including blood coagulation, wound healing, digestion and immune response.<br />

Furthermore, many studies indicate that these proteases contribute to a number of pathological<br />

conditions and play a significant role in tumour growth, invasion and metastasis [33;34].<br />

Serine proteases are characterized by the presence of a serine residue in the active site of the<br />

catalytic domain. In the active site resides the catalytic triad that is preserved in all serine<br />

proteases. This catalytic triad is composed of three amino acids; a histidine, a serine and an<br />

aspartic acid, which are essential for the catalytic ability of the protease. The specificity of the<br />

protease is determined by the size, shape and charge of the active site cleft [33].<br />

The rate of discovery of new (serine) proteases was greatly aided and accelerated by the<br />

publishing of the mouse and human genome sequences and EST databases. This led to the<br />

discovery of the hitherto unknown large family of membrane-associated serine proteases; the<br />

membrane-anchored serine proteases. Orthologues are found in all vertebrates and in humans 20<br />

members have been identified so far as presented in fig. 1 [35;36].<br />

Membrane-anchored serine proteases are tethered to the membrane either via a C-terminal<br />

transmembrane domains (Type I), a GPI-anchor or via an N-terminal transmembrane domain<br />

(Type II) as illustrated in fig. 1. The common features of Type II transmembrane serine proteases<br />

(TTSP) are a short cytoplasmic anchor, an N-terminal transmembrane domain, a C-terminal<br />

extracellular serine protease domain of the trypsin (S1) fold, and a variable length stem region<br />

containing modular domains linking the catalytic and the transmembrane domains. This stem<br />

region contains an assortment of 1-11 protein domains of six different types [3;35].<br />

14


Fig 1. Overview and domain structure of the human membrane-anchored serine proteases.<br />

Prostasin and testisin is anchored to the membrane by a GPI-anchor, whereas tryptase γ1 is<br />

anchored to the membrane through a C-terminal transmembrane domain (type I transmembrane<br />

protein). The remaining membrane-anchored serine proteases presented here are type II<br />

transmembrane proteins and are attached to the membrane by a signal anchor (SA) located close<br />

to the N terminus with a cytoplasmic extension. Type II transmembrane serine proteases are<br />

phylogenetically divided into four subfamilies; the largest being the human airway trypsin-like<br />

(HAT)/differentially expressed in squamous cell carcinoma gene (DESC) subfamily, which comprises<br />

HAT, DESC1, TMPRSS11A, HAT-like 4, and HAT-like 5 (blue shading); the hepsin/transmembrane<br />

protease, serine (TMPRSS) subfamily, which comprises hepsin, TMPRSS2, TMPRSS3, TMPRSS4,<br />

mosaic serine protease large-form (MSPL), spinesin, and enteropeptidase (yellow shading); the<br />

matriptase subfamily, which consists of matriptase, matriptase-2, matriptase-3, and polyserase-1<br />

(green shading); and the corin subfamily, which contains only corin (purple shading). Beside a<br />

serine protease domain (SPD) of the S1 fold, the modular structure of the TTSP can contain an<br />

assortment of different domains that include sea urchin sperm protein, enteropeptidase, and agrin<br />

(SEA); group A scavenger receptor (scavenger); low-density lipoprotein receptor class A (LDLR);<br />

Cls/Clr, urchin embryonic growth factor, and bone morphogenic protein-1 (CUB); meprin, A5<br />

antigen, and receptor protein phosphatase μ (MAM); and frizzled. The figure is from [3].<br />

15


This complex modular structure of most TTSPs is remarkably different from other trypsin-like<br />

serine proteases. The multi-domain structure provides the TTSPs with the capacity to interact<br />

with several interaction partners and possibly a mean for regulation of proteolytic activity [36;37].<br />

Several TTSPs are involved in fundamental cellular and developmental processes such as<br />

morphogenesis, differentiation, epithelial permeability and cellular iron transport [3].<br />

Matriptase<br />

Matiptase was first identified in 1993 as a new gelatinolytic activity expressed by cultured breast<br />

cancer cells and later 5 different groups individually cloned the cDNA from human prostate cancer<br />

cells, human ovarian carcinoma, human breast cancer cells, murine thymic cells, and human colon<br />

mucosa [38-43]. Shortly after matriptase was purified from human milk [44]. The molecular<br />

cloning of the two closely related proteases matriptase-2 and matriptase-3 was reported later<br />

[45;46]. Matriptase is also referred to as MT-SP1, TADG-15, PRSS14, TMPRSS14, SNC19, cleaving<br />

activating protease (CAP)-3, prostamin, epithin (mouse ortholog) and the gene designation is<br />

suppression of tumorigenicity 14 (ST14). For simplicity, I will use matriptase also when discussing<br />

the mouse orthologue.<br />

Matriptase expression and structure<br />

Orthologues of matriptase has been found in many vertebrates but unlike most other TTSPs,<br />

matriptase is widely expressed in epithelial compartments of many embryonic and adult tissues<br />

[10;11]. Expression analysis has been done on both human and murine tissues. mRNA and protein<br />

analyses show that matriptase is expressed in all types of epithelium, including columnar,<br />

pseudostratified, cubiodal, and squamous in humans and locates to the basolateral plasma<br />

membrane [11]. Using enzymatic gene trapping with -galactosidase as a reporter in mice showed<br />

an expression pattern of matriptase in mice overall identical to what was found in humans [12].<br />

The high degree of similarity between expression profiles of different species suggests conserved<br />

function(s) of matriptase.<br />

Full-length human matriptase is an 855 amino acid (aa) glycoprotein that lacks a classical signal<br />

peptide of app. 95 kDa. Instead the N-terminal signal anchor, which is not removed during<br />

synthesis, functions as a single-span transmembrane domain that orientates the protease in the<br />

plasma membrane as a type II integral membrane protein (fig. 1 and 2). Matriptase is a mosaic<br />

protein and is composed of a short N-terminal tail (residues 1-54) followed by the transmembrane<br />

domain, a sea urchin sperm protein, enterokinase, agrin (SEA) domain (residues 85-193), two<br />

C1r/s, urchin embryonic growth factor and bone morphogenetic protein 1 (CUB) domains<br />

(residues 214-334 and 340-447), four low-density lipoprotein receptor class A (LDLR) domains<br />

(residues 452-486, 487-523, 524-561, and 566-604) and the C-terminal trypsin-fold S1 serine<br />

protease domain (SPD; residues 614-855) with the catalytic triad composed of the residues<br />

Ser805, Asp711, and His656 [47]. Serine proteases of the S1 fold including matriptase display a<br />

pronounced specificity for the positively charged amino acids arginine and lysine at the P1<br />

position [39;46;48-50].<br />

16


Fig 2. Schematic presentation of matriptase.<br />

Matriptase consists of a transmembrane signal anchor domain, a SEA domain; two CUB domains;<br />

4 LDLR domains; and a class S1 serine protease domain.<br />

The extended substrate specificity is determined by the topology of the binding pocket and the<br />

extended specificity profile of matriptase was determined with the solution of the crystal<br />

structure of the serine protease domain of matriptase and by using positional scanning synthetic<br />

combinatorial library (PS-SCL) and substrate phage library. By these methods, the preferred<br />

cleavage sequences (P 4 -P 3 -P 2 -P 1 P 1´) of matriptase was identified to be R/K-XSR A and X-R/K-<br />

SR A, where X is a non-basic amino acid [50-52]. Thus unlike trypsin, matriptase does not<br />

indiscriminately cleave peptide substrates after Lys or Arg but requires recognition of additional<br />

residues. The extended substrate profile correlates very well with the activation motif of<br />

matriptase itself (RQAR VVGG) suggesting that matriptase can be activated by a transactivating<br />

mechanism.<br />

In vitro and cell culture studies have identified several possible substrates of matriptase; prohepatocyte<br />

growth factor/scatter factor (HGF/SF) [53;54], pro-macrophage-stimulating protein 1<br />

(MSP-1) [55], pro-urokinase-type plasminogen activator (uPA) [51;53;56], protease activated<br />

receptor 2 (PAR-2) [51;57], matrix metalloproteinases 3 (MMP-3) [58], pro-kallikreins [59],<br />

immunization M(+)HEp3 associated 135 kDa protein/CUB domain containing protein 1<br />

(SIMA135/CDCP1) [60], insulin-like growth factor binding protein-related protein 1 (IGFBP-rP1)<br />

[61], vascular endothelial growth factor receptor 2 (VEGFR-2) [62], platelet-derived growth factor<br />

(PDGF) [63;64], epidermal growth factor receptor (EGFR) [65;66], fibronectin [26], laminin [26;67]<br />

and gelatine [39].<br />

The function of the intracellular domain is unclear, though data suggest a function in interaction<br />

with the cytoskeleton to regulate localization of matriptase on the cell surface to micro domains<br />

of the plasma membrane through interaction with the actin-associated protein filamin [30;68].<br />

The extracellular non-catalytic domains are likely to function in protein-protein interactions that<br />

modulate localization, activation and inhibition and possibly substrate specificity [28;29;35]. E.g.<br />

in the complement protease cascade, CUB domains are important for the formation of C1r/C1s<br />

tetramers prior to binding of C1q and activation of the C1r and C1s proteases within the complex<br />

[69].<br />

17


Regulation of matriptase<br />

Under physiological conditions proteases are strictly regulated in time and space, thereby<br />

restricting their activity to specific subcellular sites and limiting their access to substrates. This<br />

regulation is achieved by controlled zymogen activation and interaction with inhibitors [2].<br />

Matriptase inhibitors<br />

The serpin family is the largest family of serine protease inhibitors, and complexes between<br />

matriptase and the serpins anti-thrombin III, α1-antitrypsin, and α2-antiplasmin have recently<br />

been purified from human milk and epithelial cell lines, and protease nexin 1 (PN1) complexes<br />

with matriptase was reported in vitro [70-74]. However, the physiological relevance of serpinmediated<br />

inhibition of matriptase has not yet been established. Instead, two members of the<br />

family of Kunitz-type serine protease inhibitors have been identified as essential inhibitors in<br />

regulation of matriptase.<br />

The Kunitz-type inhibitors are a class of serine protease inhibitors that make reversible<br />

interactions with their target proteases [31;75;76]. They are members of the type I<br />

transmembrane protein family and characterized by having one or two extracellular Kunitz-type<br />

serine protease inhibitor domains (KD), a single transmembrane spanning domain and a short C-<br />

terminal cytoplasmic domain [75;76]. They have an N-terminal cleavable signal sequence and a<br />

stop-transfer sequence that halts further translocation through the ER membrane and acts as a<br />

transmembrane anchor [77]. The extracellular part of the membrane bound protein can be<br />

released from the plasma membrane by ectodomain shedding (supplementary I [27]; [75]. While,<br />

HAI-1 has been purified in a complex with matriptase from human milk, seminal fluid and urine<br />

[44;78], complex formation between HAI-2 and matriptase has so far only been observed in vitro<br />

[79].<br />

HAI-1 and HAI-2 were originally identified as potent inhibitors of hepatocyte growth factor<br />

activator (HGFA) from the conditioned medium of the human stomach carcinoma cell line,<br />

MKN45 [80;81].<br />

HAI-1 is the most well studied inhibitor of matriptase and is ubiquitous expressed in epithelia and<br />

predominately localizes to the basolateral surface of epithelial cells, especially the simple<br />

columnar epithelium of ducts, tubules and mucosal surfaces [10;24]. HAI-1 is also expressed in the<br />

endothelial cells of capillaries, venules and lymph vessels [82].<br />

Tissue distribution of HAI-2 has been assessed in both human and mouse. In situ hybridisation of<br />

adult human samples showed a distribution of the inhibitor in epithelial layers of e.g. placenta,<br />

pancreas, kidney and prostate [81]. Tissue distribution of HAI-2 in mice was determined using<br />

enzymatic gene trapping and -galactosidase as a reporter system, and showed a co-localization<br />

of HAI-2 with HAI-1 and matriptase in most epithelial tissue [79]. Additionally, HAI-2 and HAI-2 are<br />

also co-expressed in developing neural tube [79;83].<br />

Human HAI-1 is composed of 513 amino acids with a 35 amino acid signal peptide, thus, the<br />

mature full-length form is composed of 478 amino acids. Full-length HAI-1 has a calculated<br />

molecular weight of 53,319 Da [29;84] and has three potential N-glycosylation sites at Asn66,<br />

Asn235 and Asn507 [80]. HAI-1 is composed of two extracellular Kunitz domains, KD1 (residues<br />

250-300) and KD2 (residues 375-425), an extracellular LDLR domain (residues 319-353), a motif at<br />

18


the N-terminal containing seven cysteines (MANSC) (residues 57-147), a transmembrane domain<br />

(residues 450-472) and a short C-terminal cytoplasmic domain as outlined in fig. 3 [80].<br />

Fig. 3. Schematic presentation of HAI-1.<br />

Domain structure of HAI-1 consists of a MANSC domain, two Kunitz-type domains separated by a<br />

LDLR domain and a transmembrane domain. HAI-1 is a type I membrane-anchored protein.<br />

The overall topology of HAI-2 is similar to that of HAI-1. Human HAI-2 is composed of 252 amino<br />

acids with a 27 amino acid signal peptide, thus, the mature full-length form is composed of 225<br />

amino acids with a calculated molecular mass of 25,415 Da [81]. HAI-2 is composed of two<br />

extracellular Kunitz domains, KD1 (residues 38-88) and KD2 (residues 133-183), and a short C-<br />

terminal region (24 aa) as outlined in fig. 4. HAI-2 has two putative N-glycosylation sites at Asn57<br />

and Asn94 [81].<br />

Fig. 4. Schematic presentation of HAI-2.<br />

Domain structure of HAI-2 consists of a transmembrane domain and two Kunitz-type domains.<br />

HAI-2 is a type I membrane-anchored protein.<br />

The amino acid in the P1 position of the reactive site is of essential importance for the inhibitory<br />

activity of KD [85]. For HAI-1 the essential amino acids in the P1 positions are arginine and lysine<br />

for KD1 and KD2, respectively, and for HAI-2 arginine is present in P1 of both KD1 and KD2<br />

[86;87]. Site-directed mutagenesis studies suggest that KD1 of both HAI-1 and HAI-2 is responsible<br />

for the inhibitory activity against proteases, and for HAI-1 this is also shown for matriptase<br />

[84;87;88].<br />

19


Matriptase proteolytic processing and activation<br />

Matriptase is synthesized as an inactive single-chain zymogen and its activation requires two<br />

sequential proteolytic processing events. The pro-enzyme is first processed at the N-termini of the<br />

stem domain after Gly149 within a conserved G SVIA motif of the SEA domain. This initial<br />

cleavage has been proposed to occur by non-enzymatic hydrolysis of the peptide bond, as it has<br />

been observed for other SEA containing proteins [89]. The second and actual activation site<br />

cleavage event occurs after Arg614 within the highly conserved R VVGG activation motif and is<br />

dependent on the first cleavage at Gly149 [28]. Hereby the disulfide-linked proteolytic active<br />

matriptase is formed (fig. 5). In spite of these processing events, the catalytic domain still remains<br />

attached to the plasma membrane due to the disulfide bond linking it to the stem domain, and<br />

strong hydrophobic interactions within the SEA domain [28;39;51;90]. It is the activity of the<br />

mature enzyme that is believed to be responsible for both the physiological and pathological<br />

functions of the enzyme identified by studies of animal models and human genetics.<br />

For the greater part of serine proteases the activation cleavage depends on another upstream<br />

active protease(s), however evidence indicate that matriptase undergoes autoactivation both in<br />

solution and in the membrane-bound form caused by weak inherent activity of the matriptase<br />

zymogen [28;29;40;91]. First off, the resemblance of the preferred cleavage sequences to the<br />

activation site motif of matriptase [39;50]. Second and most importantly, mutations of any of the<br />

amino acids in the catalytic triad of matriptase (Ser805, Asp711, or His656) result in matriptase<br />

mutants that are processed at Gly149, but unable to undergo the final proteolytic processing at<br />

Arg614. This led to a proposed transactivation mechanism for activation of matriptase [28].<br />

Autoactivation, by which two or more zymogen proteases interact and activate one another by<br />

virtue of their weak intrinsic proteolytic activity, is believed to be relevant for proteases at the<br />

apex of a protease cascades as the active site triad of serine proteases is preformed in the<br />

zymogen protease [92-94]. A well studied example of this mechanism is the activation of<br />

complement C1r protease zymogen where complex formation induce intramolecular<br />

conformational changes that result in activation by a transactivation mechanism [94].<br />

Autoactivation has also been proposed as a mechanism of activation for several other members<br />

of the TTSP family including matriptase-2 and hepsin [45;93].<br />

The activation of matriptase is extraordinary complex, and far from fully understood. Evidence<br />

suggests that the transactivation mechanism for matriptase rely on additional parameters for<br />

matriptase activation to occur including its own multi-domain structure, the plasma membrane as<br />

well as HAI-1. Membrane anchorage of full-length matriptase is required but insufficient for<br />

activation as shown in studies of matriptase in simple vesicle structure implying that a higher<br />

degree of organization within the plasma membrane is indispensable for matriptase activation<br />

[29]. A series of deletion -and point mutations demonstrated that activation of matriptase<br />

requires glycosylation of the first CUB domain (Asn302) and of the catalytic domain, as well as the<br />

presence of intact LDLR domains<br />

[28;29;95]. Thus, autoactivation of matriptase may occur by interaction of two or more<br />

neighboring SEA domain processed matriptase zymogen molecules, and possibly other cofactors<br />

to induce the necessary conformational changes in the substrate binding pocket required for<br />

catalysis [28].<br />

20


Fig. 5. Proteolytic processing of matriptase and inhibition by HAI-1.<br />

Matriptase is synthesized as a 95 kDa Full-length protease that undergoes two sequential<br />

cleavages to become fully active. The first cleavage is after Gly149 in the SEA domain and the<br />

second cleavage is after Arg614 in the conserved activation site motif N-terminal to the serine<br />

protease domain. This processing results in a disulfide linked active form of matriptase. Following<br />

activation, matriptase forms a SDS-resident complex with HAI-1.<br />

After activation, matriptase is inhibited by HAI-1 within a short time rendering active matriptase<br />

little time to act on substrates, see fig. 5 [21;30;31]. Even so, cell culture studies indicate that HAI-<br />

1 also has a role for proper expression, activation and trafficking of matriptase [28;96]. However,<br />

the requirement for HAI-1 in these processes is poorly understood. In breast cancer cells that<br />

express neither matriptase nor HAI-1 endogenously, matriptase is retained in the ER/Golgi<br />

compartments unless the cells are simultaneously transfected with HAI-1. As no difficulties in<br />

trafficking of enzymatically dead matriptase (S805A) was observed in the absence of HAI-1, the<br />

inability of wild-type matriptase trafficking without HAI-1 was suggested to be caused by the<br />

potentially toxic effect of its own unopposed proteolytic activity [96]. Also matriptase activation<br />

was impaired when matriptase was co-expressed with HAI-1 mutated in its LDLR domain [28;96].<br />

Thus, HAI-1 appears at all times to be available for matriptase, which could serve to protect the<br />

cell against aberrant matriptase proteolysis.<br />

The mechanisms triggering matriptase activation remains largely unclear, although data suggest a<br />

role for a number of extracellular stimuli. The signaling sphingolipid sphingosine 1-phosphate<br />

21


(S1P) was identified as a serum component capable of inducing matriptase activation in<br />

immortalized breast epithelial cells and the steroid sex hormone androgen is able to induce a<br />

robust but more slow activation of matriptase in human prostate cancer cells [30;31;97]. In spite<br />

of the divergence in the nature of matriptase activation in different cell types, the subsequent<br />

complex formation with HAI-1 seems to be uniform for both cell types. This is supported by the<br />

identification of suramin as a universal inducer of matriptase activation [30]. Also, exposure to a<br />

mildly acidic extracellular milieu or lowering of ionic strength induces robust and rapid matriptase<br />

zymogen activation in both cell free settings and in several epithelial cell types [29;31;95;98;99].<br />

Although matriptase remains membrane-associated after proteolytic processing and activation it<br />

is clear that matriptase is shed from the plasma membrane given that the protease has been<br />

purified from human milk [44]. N-terminal sequencing of the matriptase isoforms isolated from<br />

conditioned media of an epithelial cell line showed proteolytic cleavage either after Lys189 in the<br />

SEA domain or after Lys204 in the linker region between the SEA domain and the CUB1 domain<br />

[31]. The mechanism(s) by which matriptase is released from the plasma membrane and the<br />

proteases mediating the release remain to be identified. However, cell culture studies suggest<br />

that shedding of matriptase is dependent on the protease being fully proteolytic processed and<br />

possibly complex-formation with HAI-1 [31;78;100].<br />

Prostasin<br />

Prostasin is a GPI-anchored serine protease that was first isolated from human seminal fluid, but<br />

displays a widespread tissue distribution in both mice and humans and is co-expressed with<br />

matriptase in most murine epithelial tissues [9;101;102]. Prostasin is also termed channel<br />

activating protease 1 (CAP-1), following its ability to activate epithelial sodium channel (ENaC)<br />

[103;104]. Full-length human prostasin consists of 343 aa with a 32 aa N-terminal signal peptide<br />

that is proteolytically removed in the ER. Instead prostasin is modified with a C-terminal GPIanchor,<br />

this result in a 40 kDa mature protein that consists of an extracellular serine protease<br />

domain tethered to the outer leaflet of the plasma membrane [102;105], see fig. 6. The serine<br />

protease domain contains the catalytic triad that is formed by His53, Asp102 and Ser206<br />

[102;106]. Prostasin is synthesized as an inactive zymogen but unlike matriptase, prostasin is<br />

incapable of autoactivation and thus requires the proteolytic activity of a second protease to<br />

process the zymogen into the two-chain disulfide-linked active form [105;107]. In vitro matriptase<br />

activates prostasin by proteolytic cleavage in the amino-terminal pro-peptide region of prostasin<br />

at the Arg44 in human prostasin [16;65]. Like matriptase, prostasin has a preference for poly-basic<br />

substrates which also explain the inability of prostasin to autoactivate as the activation site motif<br />

of human prostasin is PQAR ITGG [107]. Prostasin is inhibited by PN1, HAI-1 and HAI-2 like<br />

matriptase [21;108;109].<br />

Prostasin is shed from the apical plasma membrane. In prostate epithelium this secretion is<br />

depended on C-terminal processing at the conserved residue Arg322, while secretion from kidney<br />

and lung epithelial cells depends on GPI-anchor cleavage by the endogenous GPI-specific<br />

phospholipase D1 (Gpld1) [23;101].<br />

22


Fig. 6. Schematic presentation of prostasin.<br />

Prostasin is posttranslationally modified with a GPI-anchor that tethers the single serine protease<br />

domain of prostasin to the membrane.<br />

Physiological functions of matriptase and its role in pathological processes<br />

Matriptase has been shown to play a role in a number of different proteolytic processes at the cell<br />

surface; however, the mechanisms by which matriptase modulates its effect at the molecular<br />

level is still unclear. The importance of matriptase-dependent proteolysis is derived from<br />

combined knowledge from genetic disorders and transgenic animal models, as well as cell culture<br />

studies. Below I will review physiological and pathological processes in which matriptase plays a<br />

crucial role.<br />

Matriptase is crucial for epidermal integrity<br />

In humans, a few studies have been reported identifying mutations in the matriptase gene all of<br />

which lead to various types of the skin disorder ichthyosis [13-15]. The first documented patients<br />

suffer from the rare skin disease autosomal recessive ichthyosis with hypotrichosis (ARIH)<br />

characterized by congenital ichthytosis associated with frizzy hair. The syndrome is caused by<br />

homozygosity for the missense mutation G827R in the serine protease domain of matriptase. The<br />

G827R mutant form of matriptase has a reduced proteolytic activity in vitro as the mutation<br />

affects access to the active site binding cleft [17;18]. Further information about the disease was<br />

obtained by mouse models of the disorder. Xenografting of matriptase-deficient skin onto adult<br />

athymic mice and the generation of a matriptase hypomorphic mouse model both phenocopied<br />

the microscopic hallmarks of the skin from ARIH patients [17;110]. Biochemical analysis of ST14<br />

deficient murine and human epidermis revealed highly reduced prostasin activation and reduced<br />

processing of the abundant epidermal polyprotein pro-filaggrin into filaggrin [13-17;19;110].<br />

Another example of matriptase´s importance in epidermal integrity is its recently shown<br />

implication in Netherton´s syndrome [59]. This disease is a form of ichthyosis characterized by<br />

premature stratum corneum shedding resulting in direct exposure of the living surface of the<br />

epidermis to the external environment and chronic inflammation [111-113]. The genetic<br />

23


ackground of Netherton´s syndrome is a deficiency for the protease inhibitor LEKTI [113]. In a<br />

mouse model of the syndrome, matriptase initiate a run-away kallikrein proteolytic cascade in the<br />

absence of LEKTI that results in an over accumulation of processed filaggrin [59;114].<br />

Additionally, an increase in matriptase expression has been suggested as a common basis for a<br />

range of different human skin disorders [115].<br />

The matriptase-prostasin proteolytic pathway(s) in the epidermis<br />

Matriptase and prostasin are generally co-expressed in terminally differentiated cells of stratified<br />

epithelia that do not possess any proliferative capacity [9]. Studies of both human tissue samples<br />

and murine models as well as cell culture studies have revealed that both matriptase and<br />

prostasin are important factors in regulating terminal epidermal differentiation and hair follicle<br />

development [4;9;12;17;19;116]. The phenotypes observed in mice deficient of prostasin in the<br />

skin are nearly indistinguishable from matriptase deficient mice [4;19;116]. Both mouse models<br />

display compromised epidermal tight junctions and a generalized disruption of the stratum<br />

corneum architecture. The impaired epidermal barrier function is a result of perturbation of<br />

several processes that normally takes place during terminal epidermal differentiation. They<br />

include lipid matrix formation, formation of the water-impermeable cornified layer, and<br />

desquamation (shedding of corneocytes) as well as hyperproliferation of basal keratinocytes, see<br />

fig. 7. These defects result in a lack of terminal epidermal differentiation and are accompanied<br />

with an impaired skin barrier function resulting fatal dehydration and death within 48-60 hours<br />

postnatal [4;19;116].<br />

On the molecular level, matriptase and prostasin deficiency results in greatly reduced proteolytic<br />

processing of pro-filaggrin [15;17;19;116;117]. Studies show that matriptase act upstream of<br />

prostasin in a matriptase-mediated proteolytic activation cascade most likely by directly activating<br />

the prostasin zymogen, which can be seen in fig. 7 [15-17;21]. However, the specific molecular<br />

mechanism by which matriptase and prostasin facilitates profilaggrin processing is unclear. The<br />

hypothesis is supported by the lack of active prostasin in cultured primary keratinocytes from<br />

patients with loss-of-function mutations in ST14 and in the epidermis of matriptase deficient<br />

mice, whereas active prostasin can be readily detected in wild-type human keratinocytes and in<br />

murine epidermis [4;15;16;21;116]. This correlates with activation of prostasin being suppressed<br />

by matriptase ablation in cultured human keratinocytes and that matriptase is able to activate<br />

prostasin in vitro [16;21]. Moreover, the two proteases display synchronized developmental onset<br />

of expression which correlates with acquisition of epidermal barrier function in mice [16].<br />

Together this suggests that in the skin matriptase and prostasin are functioning in the same<br />

pathway and that matriptase act upstream of prostasin in this matriptase-prostasin proteolytic<br />

cascade essential for terminal epidermal differentiation. Furthermore, it has been suggested that<br />

HAI-1 plays an important role in regulating this proteolytic cascade as inhibitor complexes of HAI-<br />

1 with both matriptase and prostasin have been detected in human cultured keratinocytes<br />

induced to differentiate into an organotypic culture model of the skin [21;115].<br />

24


Fig. 7. Matriptase´s role in the epidermis<br />

(A) Matriptase expression is confined to the uppermost living layers of the interfollicular epidermis;<br />

the transitional layer (blue color, arrowhead). This is visualized with a knock-in mouse with a<br />

promoterless β-galactosidase marker gene inserted into the endogenous matriptase gene. The<br />

basal layer is visualized using a keratin-5 antibody (red color, arrow) [12]. Deficiency of either<br />

matriptase or epidermal prostasin leads to a range of epidermal defects including loss of tight<br />

junctions, lipid extrusion and impaired processing of pro-filaggrin that result in impaired epidermal<br />

barrier function (compare B and C). Figure is from [118].<br />

Matriptase and prostasin also co-localize in a number of epithelia other than the epidermis<br />

suggestive of a role for a matriptase-prostasin proteolytic cascade in other organs as well [9].<br />

However, although we know that matriptase and prostasin are linked, we have very restricted<br />

knowledge of the pathway(s) in which they function. Identification of components acting<br />

upstream or downstream of matriptase and prostasin would greatly add to our understanding of<br />

the role(s) of these proteases in vivo.<br />

Matriptase has also been proposed to be at the apex of a matriptase-prostasin proteolytic<br />

cascade activating ENaC that is important for maintenance of salt and water homeostasis by<br />

reabsorption of Na + -ions [119;120]. This channel is also important for normal terminal<br />

differentiation of the epidermis [121]. Both matriptase and prostasin are able to activate ENaC in<br />

Xenopus oocytes [103;120]. So far evidence for a physiological role in channel activation is limited<br />

to prostasin, where fluid clearance from the lungs are impaired in mice with alveolar epitheliumspecific<br />

ablation of prostasin [122]. Intriguingly, the catalytic activity of prostasin is dispensable<br />

for its capability to activate ENaC in the Xenopus model system. Catalytically inactive prostasin<br />

induce both ENaC cleavage and activation, nonetheless cell surface expression of prostasin is still<br />

essential for activity [123;124].<br />

Matriptase has an important role in epithelial homeostasis<br />

The generation of ST14 (the gene encoding matriptase) deficient mouse models established that<br />

matriptase has a critical role in the development and maintenance of multiple epithelial tissues<br />

25


consistent with its widespread epithelial distribution [4;5;12]. Embryonic tissue specific ablation<br />

or acute ablation of matriptase in adult tissues of mice results in severe organ dysfunction and<br />

widespread epithelial demise, often associated with increased paracellular permeability, loss of<br />

tight junction function and mislocation of tight junction-associated proteins, see fig. 8. Ultimately<br />

these defects cause a reduction in body weight and ultimately death [5]. Matriptase deficiency in<br />

mice also affects hair follicle development and results in dramatically increased thymocyte<br />

apoptosis, and depletion of thymocytes [4;12].<br />

Fig. 8. Matriptase´s role in the intestinal epithelia<br />

(A) Matriptase is expressed in goblet cells (arrow) and in surface mucosal cells (arrowhead) in the<br />

murine intestine. Sustained matriptase expression is essential for maintenance of the intestinal<br />

epithelia. Ablation of matriptase disrupts tight junctions and cell polarity (compare B to C) and<br />

causes architectural distortion and compromised barrier function in the large intestine resulting in<br />

edema and diarrhea leading to premature death in mice. Figure is from [118].<br />

Matriptase is proposed to promote intestinal barrier recovery in injured intestinal mucosa of<br />

inflammatory bowel disease (IBD). Studies of human specimens and mouse models show that<br />

matriptase is down-regulated in inflamed colonic tissues from IBD patients and that matriptase<br />

has a protective role against colitis in mice [8;125]. Thus, matriptase has an essential role in<br />

development as well as maintenance of multiple types of epithelia.<br />

Importance of matriptase regulation in vivo<br />

Under normal physiological conditions, proteases are strictly regulated at the protein level.<br />

Synthesis of proteases as inactive zymogens and complex formation with protease inhibitors limit<br />

their accessibility to substrates. The importance of protease inhibitors to govern protease activity<br />

is obvious, as the activation zymogens to active proteases is an irreversible action. Therefore,<br />

important knowledge of proteases can be obtained indirectly by knock-out studies of their<br />

respective inhibitors. This section presents important knowledge about matriptase obtained by<br />

studies of knock-out animal models of its physiological inhibitors, HAI-1 and HAI-2.<br />

26


Even though matriptase is widely expressed in both embryonic and adult epithelial tissues the<br />

protease does not have critical non-redundant functions until after birth [4]. Surprisingly, the<br />

proteolytic activity of matriptase must nevertheless be under strict control during embryogenesis<br />

for correct placental development to take place. Ablation of HAI-1, HAI-2, or the combined<br />

haploinsufficiency for HAI-1 and HAI-2 leads to failure of placental labyrinth formation [15;126-<br />

128]. Correct placental development can be restored with simultaneous ablation of matriptase,<br />

indicating that the placental defects in HAI-1-deficient, HAI-2-deficient or HAI-1/HAI-2<br />

haploinsufficient mice are caused solely by unopposed matriptase activity [83;126-128]. Thus<br />

although matriptase is completely dispensable for embryogenesis, its activity needs to be strictly<br />

regulated by HAI-1 and HAI-2 for placental development to take place.<br />

HAI-1-mediated inhibition of matriptase also has an essential role in postnatal epithelial<br />

homeostasis. The generation of a chimeric mouse model with sufficient placental HAI-1, but<br />

ablation of HAI-1 in the embryo resulted in viable offspring [20;129]. These mice suffer from<br />

various defects of the keratinized epithelium and die before adulthood. The same genotype<br />

superimposed on hypomorphic expression of matriptase completely eliminates epidermal barrier<br />

and hair follicle defects in HAI-1 deficient mice [20;129]. Similar outcome is reported for<br />

zebrafish, where deletions of the HAI-1 orthologues Hai-1a and Hai-1b result in skin inflammation<br />

and a compromised epithelial integrity caused by detrimental matriptase activity that result in<br />

death app. 24 hrs after fertilization [22].<br />

Excess local matriptase activity is detrimental not only to mouse placental labyrinth formation but<br />

also for neural tube closure, which became evident with the generation of mice deficient for HAI-<br />

2. HAI-2 and matriptase are expressed in the neural and non-neural ectoderm precisely at the<br />

interface where the two neural folds fuse. Simultaneously ablation of matriptase however, only<br />

partially rescues this HAI-2 knock-out phenotype suggestive of several in vivo target of HAI-2 in<br />

neural tube closure [83]. In humans, a range of mutations in the spint2 gene encoding HAI-2 have<br />

been reported. These patients present a wide range of developmental abnormalities including<br />

duplication of internal organs and digits, craniofacial dysmorphisms, as well as anal and choanal<br />

atresia [130]. It remains to be established if these defects are caused by excess matriptase activity<br />

or unregulated activity of other serine proteases.<br />

Hence, strict and proper regulation of matriptase is important for postnatal survival, development<br />

and maintenance of several epithelia and partially for neural tube closure as these defects were<br />

caused by unregulated matriptase activity.<br />

Matriptase in carcinogenesis<br />

85 % of all cancers originate from epithelial cells that line the internal and external surfaces of the<br />

body. During transformation of normal epithelia into cancerous tissue a number of events occur;<br />

function-altering mutations of oncogenes and tumor suppressor genes, loss of epithelial cell<br />

polarity, concomitant tissue disorganization, penetration of the basement membrane and<br />

invasion of transformed epithelial cells into the underlying stroma [131]. Unlike most proteases<br />

involved in carcinogenesis which are expressed by the connective tissue supporting the epithelia,<br />

matriptase is expressed by the transformed epithelial cells themselves [6]. Matriptase is<br />

undoubtedly involved in carcinogenesis, however at present its role remains obscure. Knowledge<br />

27


of matriptase´s role in cancer comes from studies of animal models, tissue samples and cancer<br />

cell lines.<br />

The oncogenic potential of matriptase was directly shown in transgenic mice, where a modest<br />

over-expression of wild-type matriptase in the skin caused spontaneous squamous cell carcinoma<br />

formation to occur. Moreover, matriptase supports both ras-dependent and independent<br />

carcinogenesis and potentiate the effects of genotoxic exposure [6]. Matriptase expression<br />

undergoes a dramatically spatial redistribution during the transition of epidermal lesions from<br />

hyperplasia to dysplasia and becomes present in the proliferating basal compartment. Hereby<br />

matriptase is able to mediate c-Met induced activation of the PI3K-Akt-mTor pathway through<br />

matriptase-catalysed HGF activation and binding of the HGF ligand to the c-Met receptor [6;7;12].<br />

Simultaneously ablation of cMet or a corresponding over-expression of matriptase´s cognate<br />

inhibitor HAI-1 negated the oncogene potential of matriptase over-expression indicating the<br />

impact of unopposed protease activity in malignant transformation [6;7]. In addition, matriptase<br />

has been shown in vitro to activate several molecules associated with cancer progression<br />

including pro-HGF/SF [53], pro-MSP-1 [55], pro-uPA [51;53;56], PAR-2 [51;57], MMP-3 [58],<br />

SIMA135/CDCP1 [60], IGFBP-rP1 [61], VEGFR-2 [62], PDGF [63;64] and EGFR [65;66].<br />

Conversely, a recent study assigns matriptase with a role as tumor suppressor in a murine model<br />

of colitis-associated colon cancer. Intestinal-specific epithelial ablation of matriptase resulted in<br />

intrinsic intestinal permeability barrier perturbations that progressed into chronic inflammation<br />

and subsequently formation of colon adenocarcinoma highlighting the importance of matriptase<br />

in epithelial tight junction formation [8].<br />

Multiple studies have also assessed the mRNA and protein levels of matriptase during human<br />

carcinogenesis and the potential value of matriptase as a novel prognostic marker, a predictor of<br />

patient outcome and as a possible therapeutic target for human cancers. However, the findings<br />

from these studies are ambiguous.<br />

For some cancers, an increase in matriptase mRNA or protein expression levels has been<br />

associated with tumor progression in e.g. breast, cervical, ovarian and prostate cancer [132-139].<br />

In contrast, other studies report a significant downregulation of matriptase expression levels in<br />

gastric, colorectal and breast cancer [140-142]. In fact, one of the groups first to discover<br />

matriptase did so by identifying genes down-regulated in human colorectal carcinomas, hence<br />

matriptase gene annotation; suppressor of tumorigenicity-14 [41;143].<br />

Explanation for these discrepancies could be explained by different roles of matriptase in different<br />

epithelia and cancer types and partly by differences in tissue sampling, tissue composition, tumor<br />

staging and differences in quantification methodology. Expression studies including both<br />

matriptase and HAI-1 may be more informative as HAI-1 has important roles in matriptasemediated<br />

carcinogenesis in mice and for epithelial integrity in general [6;22;128;144]. An<br />

imbalance in the matriptase-HAI-1 ratio resulting in a larger proportion of free active matriptase<br />

could lead to harmful overactivation of matriptase-mediated pathways contributing to<br />

carcinogenesis [145]. This proposal of an imbalance in the matiptase-HAI-1 ratio is supported by<br />

studies in gastrointestinal, colorectal, and ovarian cancers where the protease-inhibitor ratio was<br />

shifted in favor of matriptase when comparing advanced stage tumors with low-stage lesions or<br />

corresponding tissue from control individuals [140;141;145;146].<br />

28


Technical considerations<br />

This chapter includes methodological considerations for the main methods used for the work<br />

presented in three manuscripts; paper I, II and III.<br />

Cell system<br />

Matriptase, prostasin and HAI-1 are all expressed in polarized epithelia, yet matriptase and HAI-1<br />

are located to the basolateral plasma membrane at steady state and prostasin locates to the<br />

apical plasma membrane at steady state [9;11;24;102]. However, the organization of the<br />

epithelium makes it difficult to study the polarity of the cells and the intracellular trafficking of<br />

proteins since the basolateral plasma membrane cannot be accessed. To obtain this kind of<br />

information it is applicable to utilize model systems of the polarized epithelia. Epithelial cell lines<br />

have proven to be useful tools for studying epithelium.<br />

We have previously used the Mardin Darby canine kidney (MDCK) cell line as a model system for<br />

polarized epithelia to delineate the intracellular trafficking of HAI-1 (supplementary I; [27]). In the<br />

present study (paper I and III), we decided to use the well known Caco-2 cell line as a model<br />

system for the polarized epithelium. MDCK and Caco-2 cells have both been extensively used as<br />

models of the polarized epithelium [147;148]. However combined with the available antibodies,<br />

the Caco-2 cell line offers the advantage that we can examine endogenously expressed<br />

matriptase, prostasin and HAI-1, which eliminates unwanted effects by recombinant<br />

overexpression.<br />

Caco-2 cells originate from a human colon carcinoma, yet in culture these cells spontaneously<br />

differentiate and form a polarized monolayer of cells that morphologically and functionally<br />

resemble the mature enterocytes of the small intestine. For this reason, the Caco-2 cell line has<br />

been extensively used as a model of the polarized epithelium [147]. When cultured onto<br />

Transwell filters, the apical and basolateral plasma membrane, divided by formation of tight<br />

junctions, can be accessed separately, which allows for investigations of intracellular trafficking of<br />

proteins (fig. 9).<br />

Thus, Caco-2 cells grown on Transwell filters are a good model system for studying the<br />

intracellular transport of matriptase, prostasin and their common inhibitor HAI-1. In our studies<br />

(paper I and III) we have used 11 days post-confluent Caco-2 cells grown on Transwell filters as<br />

the level of HAI-1 complexed matriptase reaches a plateau at this differentiation level<br />

(unpublished results, Stine Friis).<br />

Tight junction formation and barrier function of Transwell filter grown Caco-2 cells can be<br />

assessed in several ways. It can be measured by transepithelial electrical resistance (TEER) also;<br />

tracer molecules can be used to measure the permeability of the Caco-2 monolayer, e.g. lucifer<br />

yellow or phenol red [149;150]. However, a much simpler way of evaluating barrier function of<br />

cells grown on Transwell filters is to assess whether the monolayer of cells is able to maintain a<br />

difference in medium level between the inner and outer chamber over night [27].<br />

29


Fig. 9. Monolayers of polarized Caco-2 cells on porous filters.<br />

Caco-2 cells spontaneously form a polarized monolayer of cells when grown on Transwell filters. At<br />

confluence, the cells form tight junctions that separate the apical and basolateral extracellular<br />

spaces. This setting allows for separate access to the apical and basolateral plasma membrane<br />

domains and media. The figure shows a cross section of the Transwell filter insert in the culture<br />

dish.<br />

Although cell lines are extremely useful for obtaining new knowledge about molecular mechanism<br />

responsible for different phenomena, data from one cell line is not always comparable with data<br />

from other cell line nor can one expect data to be transferable to whole organism settings. One<br />

reason for this is the immortalized feature of most cell lines. However, cell lines make good model<br />

systems to study important biological pathways.<br />

Protease pull down assays<br />

Variations over protease inhibitor mediated pull down were applied in paper I and II. These<br />

techniques were used to identify activation state and to assay for the presence of active protease.<br />

Different set ups were used in the two studies but they rely on the same concept.<br />

Co-immunoprecipitation (co-IP) is a well established method for small scale purification of<br />

antigens and their interaction partners and allows for identification of activation status of<br />

proteins. In a co-IP, the target antigen has the function of bait and is used to co-precipitate a<br />

binding partner/protein complex (prey) from a crude cell lysate or tissue extract. An antibody<br />

against the bait is incubated with a lysate either as pre-immobilized onto an insoluble support or<br />

as a free un-bound antibody in combination with an insoluble support e.g. protein G sepharose as<br />

outlined in fig. 10. The main disadvantage of free antibody protocol for immunoprecipitation and<br />

co-IP is that the conditions used to elute the precipitated antigen also release the antibody.<br />

Depending on the size of the antigen, the heavy and light chains of the antibody can completely<br />

30


mask the detection of the antigen/interaction partners in Western blot analysis. However this is<br />

circumvented by cross-linking the antibody to the resin.<br />

Fig. 10. Immunoprecipitation<br />

A suitable antibody is added to a cell lysate or a tissue extract either pre-immobilized onto an<br />

insoluble support or as a free unbound antibody to bind the protein of interest. Gentle incubation<br />

allows the antigen to by immobilized and purified by means of the antibody and the insoluble<br />

support. In the pre-immobilized antibody approach the antibody has been cross-linked to the<br />

insoluble support. Immunoprecipitation of intact protein complexes is known as co-IP.<br />

Immunoprecipitated proteins and their binding partner(s) can be detected by SDS-PAGE and<br />

Western blot analysis. Figure is from [151].<br />

31


We applied co-IP to determine the presence activated matriptase and activated prostasin in<br />

extracts of murine placentas in paper II, as simple detection of the activated proteases in lysates<br />

of placentas was unsuccessful by direct Western blotting. We used a free antibody protocol (the<br />

latter approach) to co-IP matriptase and prostasin with a HAI-1 antibody and used protein G<br />

sepharose that binds to the Fc region of immunoglobulin G as the insoluble support. As unspecific<br />

binding to the protease G sepharose can occur, the lysates were pre-incubated with protein G<br />

sepharose alone to minimize unspecific binding in the actual co-IP reaction. Placentas from<br />

matriptase deficient and prostasin deficient embryos were used as negative controls, respectively.<br />

Using co-IP it is often anticipated that associated proteins (prey) is related to the function of bait<br />

protein. However, this is merely an assumption that needs further verification. Though, in this<br />

case the relationship between matriptase, prostasin and HAI-1 is already well-established [16;21].<br />

Thus, we can use the technique to identify the presence matriptase and prostasin activation in<br />

placental extracts from embryos of different genotype.<br />

In paper I, we use a derivation of immunoprecipitation to detect the presence of active matriptase<br />

and prostasin. Instead of using antibodies to immobilize matriptase and prostasin, we applied the<br />

general serine protease inhibitors aprotenin and leupeptin assuming that only active serine<br />

proteases and not their zymogen counterparts will be able to bind to these serine protease<br />

inhibitors. As HAI-1 forms reversible inhibitor complexes with proteases, immobilized aprotenin<br />

and leupeptin could potentially compete with HAI-1 to bind matriptase and prostasin. This can be<br />

ruled out by a complicated experimental setting where the inhibitor-coupled pull down is<br />

performed on a sample in which matriptase-HAI-1 complex formation is induced. The major<br />

disadvantage of this method is that detection of active matriptase occurs in solution which may<br />

deviate from live cell culture settings.<br />

Biotinylation assays<br />

The highly specific interaction of biotin with avidin/streptavidin is a useful tool in designing<br />

nonradioactive purification and detection systems. The extraordinary affinity of avidin<br />

/streptavidin for biotin is the strongest known noncovalent interaction of a protein to a ligand and<br />

allows biotin-containing molecules in a complex mixture to be isolated by exploiting this highly<br />

stable interaction. For paper I and III included in this thesis, we have exploited the advances of the<br />

biotin-avidin interaction to study matriptase, prostasin and HAI-1. In paper I, we have made use of<br />

biotinylation assays for studying the subcellular trafficking of matriptase, prostasin and HAI-1 in<br />

Transwell filter grown polarized Caco-2 cells. In paper III, we have exploited the<br />

streptavidin/avidin–biotin interaction to extract proteases with a peptidolytic activity towards a<br />

biotinylated chloromethyl ketone peptide inhibitor designed with a preferred substrate sequence<br />

of matriptase.<br />

32


Fig. 11. Reaction of S-NHS-SS-biotin with primary amine on proteins.<br />

Primary amines of proteins react with NHS-esters by nucleophilic attack, whereby the N-<br />

hydroxysulfosuccinimide (NHS) is released as a byproduct. The NHS ester group on this reagent<br />

reacts with the ε-amine of lysine residues and forms a stable product. Hydrolysis of the NHS-ester<br />

competes with the reaction in aqueous solution and increases with increasing pH. α-amine groups<br />

present on the N-termini of peptides also react with NHS esters but these α-amines are seldom<br />

accessible for conjugation in proteins. Figure is from [152].<br />

Biotin is a relatively small molecule and can thus be conjugated to many proteins without<br />

significant affect on the target proteins biological activities. More biotin molecules can bind to a<br />

single protein which greatly increases the sensitivity of many assay procedures. Additionally,<br />

biotin has been modified in numerous ways to accommodate particular applications. In paper I,<br />

we have utilized the membrane impermeable biotin derivative sulfosuccinimidyl-2-<br />

[biotinamido]ethyl-1,3-dithiopropionate (S-NHS-SS-biotin) because this conjugate is cell<br />

membrane impermeable due to the charged sulfonate group and is thus favorable for cell surface<br />

biotinylation. The thiol-cleavable spacer arm of S-NHS-SS-biotin harbors two important<br />

properties. First, the extended spacer arm reduces steric hindrance associated with<br />

avidin/streptavidin binding and thereby enhances the interaction. Second, the S-S cleavable<br />

spacer arm allow for reversible biotinylation, by treatment with reducing agents, which is a<br />

required feature when performing transport studies of cell surface proteins. Specifically, we use<br />

this feature to determine the endocytosis and transcytosis properties of matriptase, prostain and<br />

HAI-1. Finally, S-NHS-SS-biotin reacts with primary amines readily available on the cell surface of<br />

proteins (fig. 11).<br />

33


Fig. 12. Assay for detection of active matriptase in cell culture<br />

To establish an assay for the detection of active matriptase, we developed a synthetic peptide with<br />

binding preference for the active cleft in matriptase. This peptide is coupled to both a biotin-group<br />

and to a chloromethyl ketone (Cmk) group. When the Cmk group is in close proximity to proteases,<br />

covalent bond formation occurs by alkylation of the active site histidine. Active proteases with<br />

binding preferences for the Cmk inhibitory peptide is labeled and extracted by streptavidin pull<br />

down of the biotin group. Specificity for matriptase is obtained via Western blot analysis and<br />

matriptase specific antibodies.<br />

For paper III, we make use of the N-terminal biotin moiety of the CMK peptide inhibitor, biotin-<br />

RQRR-Cmk, as a tag to extract peptidolytic active matriptase from Caco-2 cell lysates. The peptide<br />

sequence of the tetra peptide was designed to obtain the highest specificity towards matriptase<br />

and was deduced from a preferred substrate sequence of matriptase [50]. This peptide enables us<br />

to biotin-label peptidolytic active matriptase in live and intact Caco-2 cells (fig. 12), as opposed to<br />

the inhibitor-coupled sepharose used in paper I that label active matriptase in solution.<br />

Once biotin is attached to its target molecule, the molecule can be immobilized using biotin<br />

binding molecules. In this study we have made use of streptavidin (paper I and III) and monomeric<br />

avidin (paper I). The high affinity biotin-streptavidin binding allows for extraction of biotin from<br />

very complex mixtures, and dissociation between biotin and streptavidin requires harsh<br />

denaturing conditions e.g. boiling in SDS-PAGE sample buffer. In contrast, monomeric avidin binds<br />

biotin in a reversible manner, and gentle recovery of biotinylated molecules is feasible by elution<br />

with free excess biotin. Monomeric avidin can facilitate purification of functional proteins and in<br />

paper I, we employ this in recovery of the SDS-resistant matriptase-HAI-1 complex in Western blot<br />

analysis.<br />

34


Results<br />

Data included in this <strong>PhD</strong> thesis is presented in the following three manuscripts:<br />

Manuscript I<br />

Transport via the transcytotic pathway makes prostasin available as substrate for matriptase<br />

Manuscript II<br />

Reduced prostasin (CAP1/PRSS8) activity eliminates HAI-1 and HAI-2 deficiency associated<br />

developmental defects by preventing matriptase activation<br />

Manuscript III<br />

Novel assay for detection of active matriptase<br />

35


Paper I<br />

Transport via the transcytotic pathway makes prostasin available as substrate for matriptase<br />

Stine Friis 1 , Sine Godiksen 2 , Jette Bornholdt 1 , Joanna Selzer‐Plon 1 , Hanne Borger Rasmussen 3 ,<br />

Thomas H. Bugge 4 , Chen‐Yong Lin 5 and Lotte K. Vogel 1<br />

1 Department of Cellular and Molecular Medicine, University of Copenhagen, Copenhagen,<br />

Denmark.<br />

2 Department of Biology, University of Copenhagen, Copenhagen, Denmark.<br />

3 Department of Biomedical Science, University of Copenhagen, Copenhagen, Denmark<br />

4 Proteases and Tissue Remodeling Unit, National <strong>Institut</strong>e of Dental and Craniofacial Research,<br />

National <strong>Institut</strong>es of<br />

Health, Bethesda, USA<br />

5 Department of Biochemistry and Molecular Biology, Greenebaum Cancer Centre, University of<br />

Maryland, Baltimore,<br />

USA<br />

Published in Journal of Biological Chemistry, February 2011.<br />

36


THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 286, NO. 7, pp. 5793–5802, February 18, 2011<br />

Printed in the U.S.A.<br />

Transport via the Transcytotic Pathway Makes Prostasin<br />

Available as a Substrate for Matriptase *<br />

Received for publication, September 29, 2010, and in revised form, December 1, 2010 Published, JBC Papers in Press, December 10, 2010, DOI 10.1074/jbc.M110.186874<br />

Stine Friis ‡ , Sine Godiksen § , Jette Bornholdt ‡ , Joanna Selzer-Plon ‡ , Hanne Borger Rasmussen , Thomas H. Bugge ,<br />

Chen-Yong Lin**, and Lotte K. Vogel ‡1<br />

From the Departments of ‡ Cellular and Molecular Medicine, § Biology, and Biomedical Science, University of Copenhagen,<br />

2200 Copenhagen, Denmark and the Proteases and Tissue Remodeling Unit, NIDCR, National <strong>Institut</strong>es of Health,<br />

Bethesda, Maryland 20892, and the **Department of Biochemistry and Molecular Biology, Greenebaum Cancer Center,<br />

University of Maryland, Baltimore, Maryland 21201<br />

Thematriptase-prostasinproteolyticcascadeisessentialfor<br />

epidermaltightjunctionformationandterminalepidermal<br />

differentiation.Thisproteolyticpathwaymayalsobeoperative<br />

inavarietyofotherepithelia,asbothmatriptaseandprostasin<br />

areinvolvedintightjunctionformationinepithelialmonolayers.However,inpolarizedepithelialcellsmatriptaseismainly<br />

locatedonthebasolateralplasmamembranewhereasprostasinismainlylocatedontheapicalplasmamembrane.Todeterminehowmatriptaseandprostasininteract,wemappedthe<br />

subcellularitineraryofmatriptaseandprostasininpolarized<br />

colonicepithelialcells.Weshowthatzymogenmatriptaseis<br />

activatedonthebasolateralplasmamembranewhereitisable<br />

tocleaverelevantsubstrates.Afteractivation,matriptase<br />

formsacomplexwiththecognatematriptaseinhibitor,hepatocytegrowthfactoractivatorinhibitor(HAI)-1andisefficientlyendocytosed.Themajorityofprostasinislocatedon<br />

theapicalplasmamembranealbeitaminorfractionofprostasinispresentonthebasolateralplasmamembrane.Basolateral<br />

prostasinisendocytosedandtranscytosedtotheapicalplasma<br />

membranewherealongretentiontimecausesanaccumulationofprostasin.Furthermore,weshowthatprostasinonthe<br />

basolateralmembraneisactivatedbeforeitistranscytosed.<br />

Thisstudyshowsthatmatriptaseandprostasinco-localizefor<br />

abriefperiodoftimeatthebasolateralplasmamembraneafterwhichprostasinistransportedtotheapicalmembraneas<br />

anactiveprotease.Thisstudysuggestsapossibleexplanation<br />

forhowmatriptaseorotherbasolateralserineproteasesactivateprostasinonitswaytoitsapicaldestination.<br />

Thetrypsin-likemembraneserineproteasematriptaseis<br />

essentialformaintenanceofmultipletypesofepithelia.ConditionalablationoftheSt14genecodingformatriptasein<br />

intestine,kidney,andlungofadultmiceresultsinweightloss,<br />

severedeclineinhealthanddeathwithin2weeks,causedby<br />

* This work was supported, in whole or in part, by the NIDCR Intramural<br />

Research Program and by National <strong>Institut</strong>es of Health Grant R01-CA-<br />

123223. This work was also supported by The Harboe Foundation, The<br />

Augustinus Foundation, The Brothers Hartmanns Foundation, The A.P.<br />

Møllers Foundation for the Advancement of Medical Science, The Cluster<br />

of Cell Biology at the University of Copenhagen, and the Lundbeck<br />

Foundation.<br />

1 To whom correspondence should be addressed: Blegdamsvej 3, Bldg. 6.4,<br />

2200 Copenhagen N, Denmark. Tel.: 45-35-32-77-87; Fax: 45-35-36-79-80;<br />

E-mail: vogel@sund.ku.dk.<br />

organdysfunctionassociatedwithincreasedpermeabilityand<br />

lossoftightjunctions(1).Knockdownofmatriptaseby<br />

siRNAinacellmodeloftheintestinalepitheliumcauseda<br />

leakybarrier,impairedabilitytodeveloptransepithelialelectricalresistance(TEER)<br />

2 andenhancedparacellularpermeabilitythroughregulationoftightjunctionproteins(2).Togetherthesedatasuggestakeyroleformatriptaseinepithelial<br />

barrierfunctionandtightjunctionassembly.<br />

Prostasin(alsoknownasCAP1andPRSS8)isaGPI-anchoredtrypsin-likeserineprotease.Prostasinisco-expressed<br />

withmatriptaseinmostepithelialtissuesincludingtheepidermis,kidney,andcolon(3).Prostasinproteolyticactivity<br />

hasalsobeensuggestedtopromotethedevelopmentoffunctionaltightjunctions,TEERandparacellularpermeability<br />

(4–6).Unlikematriptase,whichundergoesefficientautoactivation,theprostasinzymogenisnotabletoauto-activate,<br />

andformationofactiveprostasinrequiresactivationsite<br />

cleavagebyothertrypsin-likeserineproteases.Strongdata<br />

suggestthatmatriptaseactsupstreamofprostasininazymogencascadeintheepidermis.Thesevereepidermaldefectsof<br />

matriptasedeficiencyappeartobeaconsequenceoflackof<br />

activeprostasin.Onlytheinactiveformofprostasinisfound<br />

inmatriptase-deficientmiceandmatriptase-deficientand<br />

prostasin-deficientmicehavenearlyidenticalphenotypes<br />

withcompromisedepidermaltightjunctionformationandno<br />

terminalepidermaldifferentiation(6–11).Furthermore,it<br />

hasbeenshowninvitrothattheserineproteasedomainof<br />

matriptaseisdirectlyabletocleavethezymogen-formof<br />

prostasin,togenerateproteolyticallyactiveprostasin(11).<br />

Matriptase-dependentactivationofprostasinwasrecently<br />

demonstratedinahumanorganotypicskinmodelandin<br />

matriptase-deficienthumanepidermis(6,12).<br />

Theplasmamembraneofapolarizedepithelialcellisdividedintoanapicalandabasolateralplasmamembranedomainseparatedbytightjunctions.Thetightjunctionsprevent<br />

diffusionofmembraneproteinsbetweenthetwomembrane<br />

domains.Inpolarizedepithelialcellstherearetwopathways<br />

fornewlysynthesizedproteinstoreachtheapicalplasma<br />

membrane:ThedirectpathwayfromthetransGolginetwork<br />

directlytotheapicalplasmamembraneandtheindirectpath-<br />

2 The abbreviations used are: TEER, transepithelial electrical resistance;<br />

HAI-1, hepatocyte growth factor activator inhibitor 1; ENaC, epithelial<br />

sodium channel.<br />

Downloaded from www.jbc.org at DEF Consortium - Københavns Universitetsbibliotek, on June 15, 2011<br />

FEBRUARY 18, 2011 • VOLUME 286 • NUMBER 7 JOURNAL OF BIOLOGICAL CHEMISTRY 5793


Matriptase Activates Prostasin on the Basolateral Plasma Membrane<br />

way,wherenewlysynthesizedproteinsviathebasolateral<br />

plasmamembraneareendocytosedandtranscytosedtothe<br />

apicalplasmamembrane(13).Ithasbeenreportedthat<br />

matriptaseismainlylocatedatthebasolateralplasmamembraneinratenterocytesandotherpolarizedepithelialcells<br />

(14,15).However,proteolyticallyshedmatriptaseincomplex<br />

withHAI-1waspurifiedfromhumanmilksuggestinganapicalsecretion(16).Conversely,thematriptasesubstrate,prostasin,ismainlylocatedattheapicalplasmamembraneofpolarizedepithelialcells(17,18).<br />

Thepresentstudyaimstodeterminewhereinthepolarized<br />

epithelialcellactivematriptaseinteractswithitssubstrate<br />

prostasin,inordertoexplainhowabasolateralproteasecan<br />

cleaveandactivateanapicallylocatedsubstrate.Matriptaseis<br />

synthesizedasaninactive,single-chainzymogen.Itsactivationrequirestwosequentialendoproteolyticcleavages.The<br />

firstproteolyticprocessingcleavageoccursafterGly-149,yet<br />

theprocessedformremainstightlyassociatedwiththe<br />

membrane.<br />

Matriptaseis,subsequently(anddependentonthefirst<br />

cleavage)cleavedafterArg-614intheserineproteasedomain<br />

togainproteolyticactivity.Shortlyafteractivation,matriptase<br />

formsacomplexwithHAI-1,wherebymatriptaseisenzymaticallyinhibited.Hence,substratesshouldbepresentatthe<br />

samelocationasmatriptaseactivationtakesplace.<br />

Wepresentbiochemicaldatashowingthatamatriptaseprostasinzymogencascadeisindeedpossibleinpolarized<br />

epithelialcells.Weshowthatmatriptaseiscleavedtoitsactiveformonthebasolateralplasmamembrane,subsequently<br />

inhibitedbyHAI-1followedbyendocytosis.Importantly,we<br />

findprostasinpresentonthebasolateralplasmamembrane<br />

duringmatriptaseactivation.Furthermore,thebasolateral<br />

prostasinisactiveandtranscytosedtotheapicalplasma<br />

membranewhereitaccumulates.Theseresultsdemonstrates<br />

thatmatriptaseandprostasinmayfunctionallyinteractin<br />

polarizedepithelialcells,despitetheir,respective,basolateral<br />

andapicallocationsatsteadystate.<br />

EXPERIMENTAL PROCEDURES<br />

CellCulture—Caco-2cellsweregrowninminimalessential<br />

mediumsupplementedwith2mM L-glutamine,10%fetalbovineserum,1nonessentialaminoacids,100units/mlpenicillin,and100<br />

g/mlstreptomycin(Invitrogen)at37°Cinan<br />

atmosphereof5%CO 2 .Forallexperiments,210 6 cells<br />

wereseededinto0.4 m-pore-size24mmTranswellfilter<br />

chamber(Corning)allowingseparateaccesstotheapicaland<br />

basolateralplasmamembrane.Cellsweregrownuntilday11<br />

postconfluencebeforetheywereusedforexperiments.The<br />

tightnessoffilter-growncellswasassayedbyfillingtheinner<br />

chambertothebrimandallowingittoequilibrateovernight.<br />

Thecellculturemediumwaschangedeveryday.<br />

Biotinylation,Internalization,andBiotin-removal—Caco-2<br />

cellsgrownontranswellfilterswerewashedthreetimeswith<br />

ice-coldPBS(PBSsupplementedwith0.7mMCaCl 2 and<br />

0.25mMMgCl 2 )onbothapicalandbasolateralside.Thecells<br />

werebiotin-labeledfor30minat4°C,eitherfromtheapical<br />

orthebasolateralside,with1mg/mlEZ-link TM Sulfo-NHS-<br />

SS-Biotin(Pierce)dissolvedinPBS.Afterbiotinlabelingthe<br />

cellswerewashedtwicewithice-coldPBS.Residualbiotin<br />

wasquenchedfor5minat4°Cwith50mMglycineinPBS<br />

andthecellswerewashedagainwithPBS.Forinternalizationexperiments,preheatedmedia(serum-freeMEMmediumcontaining20mMNaHCO<br />

3 ,2mM L-glutamine,100<br />

units/mlpenicillin,and100 g/mlstreptomycin(Invitrogen))<br />

wasaddedandthecellswereincubatedat37°Ctoregainnormaltrafficking.Forendocytosisandtranscytosisexperiments,<br />

surfacebiotinwasremovedafterincubationasindicatedusingthenon-membranepermeablereducingagentglutathione<br />

(SigmaAldrich).Surfaceexposedbiotinwasremovedfrom<br />

eitherapicalorbasolateralsidewith16mg/mlglutathionein<br />

75mMNaCl,75mMNaOH,1mMEDTA,and0.5%BSAin<br />

H 2 Oundergentleagitationfor220min.Cellswerewashed<br />

withPBSandresidualglutathionewasinactivatedwith5<br />

mg/mliodoacetamide(SigmaAldrich)inPBSfor5min.A<br />

setofparallelsampleswasleftwithouttheglutathionereductiontomonitorthetotalamountofbiotinylatedproteinpresentthroughthecourseoftheexperiment.Cellswerewashed<br />

twiceinPBS,andlysedinPBScontaining1%TritonX-100,<br />

0.5%deoxycholateandproteaseinhibitors(10mg/literbenzamidine,2mg/literpepstatinA,2mg/literleupeptin,2mg/l<br />

antipain,and2mg/literchymostatin).Forinhibitor-Sepharosepull-downs,proteaseinhibitorswereomittedfromthe<br />

lysisbuffer.<br />

MonomericAvidin/StreptavidinPrecipitationofBiotinylatedProteins—Lysatesfrombiotin-labeledcellswerecentrifugedat20,000<br />

gfor20mintopellettheinsolublematerial.<br />

Thesupernatantwastransferredtocleaneppendorftubes<br />

witheitherPiercemonomericavidinagarose(Pierce)(120<br />

l/24mmfilter)orPiercestreptavidinagarose(50 l/24<br />

mmfilter),preparedasdescribedbymanufacturer.After<br />

overnightincubationat4°Cwithend-over-endrotation,the<br />

agarosewaswashedthreetimeswith50mMTris-HCl,pH<br />

6.25.Biotinylatedproteinswereelutedfromthemonomeric<br />

avidinagarosewith4mMbiotin(Pierce)inPBSfor30min<br />

followedbyadditionofSDSsamplebuffer.Forelutionfrom<br />

streptavidin-agarose,thesampleswereboiledinSDSsample<br />

buffer.<br />

WesternBlot—The2SDSsamplebuffer(125mMTris-<br />

HCl,25mMEDTA,pH6.8,4%SDS,5%glycerol,0.01%bromphenolblue)didnotcontainanyreducingagentandthesampleswerenotboiledpriortoSDS-PAGEtopreventprotein<br />

complexesfromdissociating,unlessotherwisespecified(0.1 M<br />

dithiothreitol,DTT).AlllysateswereincubatedwiththeSDS<br />

samplebufferfor10minatroomtemperaturebeforegel<br />

loading.Theproteinswereseparatedon7%acrylamidegels<br />

madeinthelaboratoryandtransferredtoImmobilon-PPVDF<br />

membranes(Millipore).Themembraneswereblockedwith<br />

10%nonfatdrymilkinPBScontaining0.1%Tween-20<br />

(PBST)for1hatroomtemperature.TheindividualPVDF<br />

membraneswereprobedwithprimaryantibodiesdilutedin<br />

1%nonfatdrymilkinPBSTat4°Covernight.Thenextday<br />

themembraneswerewashed3withPBSTandthebinding<br />

ofprimaryantibodieswasfollowedbyrecognitionwithsecondaryhorseradishperoxidase(HRP)-conjugatedsecondary<br />

antibodies(Pierce).After3washwithPBST,thesignalwas<br />

developedusingtheECLreagentSuperSignalWestFemto<br />

Downloaded from www.jbc.org at DEF Consortium - Københavns Universitetsbibliotek, on June 15, 2011<br />

5794 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 286 • NUMBER 7 • FEBRUARY 18, 2011


Matriptase Activates Prostasin on the Basolateral Plasma Membrane<br />

MaximumSensitivitySubstrate(Pierce),accordingtotheprotocolsuppliedbythemanufacturerandvisualizedwithaFuji<br />

LAS1000-camera(Fujifilm,SwedenAB).Forgraphs,thefree<br />

onlinesoftwareImageJ(createdbyWayneRasband,NIH,<br />

Bethesda)wasusedtoquantifythebandsontheblot.Values<br />

inthegraphrepresentthesumofallthebandsvisualizedon<br />

thegel.Furthermore,thegraphrepresentsthemeanofthree<br />

independentexperimentsandispresentedwiththestandard<br />

deviation.<br />

Antibodies—Theantibodiesusedweremonoclonalmouse<br />

anti-humanantibodiesM32andM69(19).TheantibodyM32<br />

detectsallformsofmatriptase,includingzymogen,activated<br />

form,andcomplexes.Undertheconditionsusedinthisstudy,<br />

theantibodyM69onlydetectedthematriptase-HAI-1complexandM69reactivematerialishereafterreferredtoasthe<br />

matriptase-HAI-1complex.Theotherantibodiesusedwere<br />

polyclonalrabbitanti-humanmatriptaseraisedagainstthe<br />

serineproteasedomainofmatriptase(Cat.no.IM1014,Calbiochem),mouseanti-humanHAI-1antibodyM19(19),and<br />

mouseanti-humanprostasinantibody(Cat.no.612173,BD<br />

TransductionLaboratories).M32,M69,andpolyclonalgoat<br />

anti-humanHAI-1(cat.no.AF1048,R&D)antibodieswere<br />

usedforimmunocytochemistry.<br />

ProteasePull-downwithProteaseInhibitor-coupledSepharose4B—Theproteaseinhibitorsaprotinin(5mg/mlSepharose),leupeptin(5mg/mlSepharose),andsoybeantrypsin<br />

inhibitor(5mg/mlSepharose)wereimmobilizedtoCNBractivatedSepharose<br />

TM 4B(GEHealthcare),asspecifiedbythe<br />

manufacturer’sinstructions.Caco-2cellswerebiotin-labeled<br />

andbiotinylatedproteinswerepulleddownwithmonomeric<br />

avidinagarose.Thebiotinylatedproteinsweregentlyeluted<br />

with4mMbiotinin50mMTris-HCl,pH8.5.Thebiotin-eluatewasseparatedfromtheavidin-agaroseandincubatedwith<br />

60 lproteaseinhibitor-coupledSepharosein50mMTris-<br />

HCl,pH8.75at37°Cfor30min.TheinhibitorSepharosewas<br />

washed3with50mMTris-HCl,pH6.5.Proteaseswere<br />

elutedfromtheinhibitor-coupledSepharoseusing0.1 Mglycine,pH2.4.Sampleswereneutralizedwith1MTrisimmediatelyafterelution.TheeluateswereaddedtoSDSsample<br />

bufferandanalyzedbyWesternblotting.<br />

GelatinZymography—Biotinylatedmonomericavidinagarose-purifiedproteinswereseparatedona7%SDS-polyacrylamidegelcontaining0.1%gelatin.Theproteinswerere-naturedbywashingthegelatingel230minin2.5%Triton<br />

X-100inH 2 O.TowashoutexcessTritonX-100,thegelwas<br />

washed210mininH 2 Oandthereafterincubatedin50mM<br />

Tris-HCl,pH8.75overnightat37°C.Gelatinolyticbandson<br />

thegelwerevisualizedbyCoomassieBrilliantBluestaining.<br />

Immunofluorescence—Caco-2cellsgrownonfilterswere<br />

fixedfor20minin4%paraformaldehydeinPBS(Bie&Berntsen)atroomtemperature.Thefollowingwasperformedat<br />

4°C.Cellswerepermeabilizedwith0.05%TritonX-100in<br />

PBSfor20min.UnspecificstainingwasblockedwithPBS<br />

containing3%BSA(PBS/BSA)for30min.CellswereincubatedwithprimaryantibodydilutedinPBS/BSAfor1.5h,<br />

washed3inPBSfollowedbyincubationwithrelevantAlexa<br />

Fluor-conjugatedsecondaryantibodies(Invitrogen)for1h.<br />

Whereindicatedthenucleiofthecellswerevisualizedby<br />

4,6-diamidino-2-phenylindole(DAPI)staining.Thecells<br />

werefinallymountedwithProlongGoldmountingmedium<br />

(Invitrogen)andsubjectedtolaserscanningconfocalmicroscopyusingtheLeicaTCSSP2systemandZeissLS700system.<br />

RESULTS<br />

Steady-stateDistributionofMatriptase,HAI-1,and<br />

Prostasin—Whereinthecellmatriptasecleavesitssubstrates<br />

isnotwellunderstood,howeveritisfundamentaltounderstandinghowmatriptasemaintainsepithelialintegrity.Traffickingandactivationofendogenousmatriptase,prostasin,<br />

andHAI-1wasthereforestudiedinCaco-2cells-ahuman<br />

colonepithelialcellline.Uponreachingconfluence,Caco-2<br />

cellsspontaneouslydifferentiateintoatightmonolayerof<br />

polarizedcellswithanapicalandabasolateralplasmamembrane,separatedbytightjunctions.Matriptaseexpression<br />

increasesduringCaco-2differentiation(2)consistentwiththe<br />

higherlevelsofmatriptaseattheintestinalvilloustip(20).<br />

Theamountofmatriptase-HAI-1complexalsoincreasesduringdifferentiationandreachesaplateauaroundday7–10in<br />

differentiatingCaco-2cells(datanotshown).Forthatreason,<br />

Caco-2cellsatday11post-confluencewereusedforthefollowingexperiments.<br />

Wefirstinvestigatedthesteadystatedistributionof<br />

matriptase,HAI-1,andprostasinbetweentheapicalandthe<br />

basolateralplasmamembranedomains.Surfacebiotinylation<br />

experimentsshowedthatthemajorityofmatriptasewaspresentamongbasolateralmembraneproteinsandmostprevalent<br />

ina70kDaformcorrespondingtotheextracellulardomainof<br />

matriptasecleavedatGly-149(Fig.1,lanes1and2),representingeitherzymogenmatriptaseorenzymaticallyactive<br />

matriptase.Asmallfractionofthebasolateralmatriptasewas<br />

detectedina120kDaform(Fig.1,lane2).The120kDaform<br />

wasidentifiedasmatriptase-HAI-1complexbydetectionwith<br />

boththeM69andanti-HAI-1antibodies(Fig.1,lanes4and<br />

6).Twomatriptaseformsbelowthe120kDacomplexwere<br />

alsodetectedonthebasolateralplasmamembrane(Fig.1,<br />

lanes2and4).Thiscouldpossiblybedegradationproductsof<br />

thematriptase-HAI-1complexoractivatedmatriptasein<br />

complexwithotherinhibitors(21).<br />

Full-lengthHAI-1hasanestimatedmolecularweightof55<br />

kDaandwasdetectedonboththeapicalandthebasolateral<br />

plasmamembrane(Fig.1,lanes5and6).Apicallylocated<br />

HAI-1displayedahighermobilitythanbasolaterallylocated<br />

HAI-1inthepresenceofSDS(Fig.1,lanes5and6)butthe<br />

samemobilitywhenthesampleswereboiled(Fig.1,lanes7<br />

and8).ThiscouldindicateaconformationaldifferencemakingapicalHAI-1moreresistanttodenaturationwithSDS.<br />

Furtherstudiesareneededtoelucidatethenatureofthetwo<br />

forms.AHAI-1complexaround85kDawasalsodetectedon<br />

thebasolateralplasmamembrane(Fig.1,lane6).An85kDa<br />

HAI-1complexwithmatriptaseaswellasprostasinhaspreviouslybeenreported(22,23).<br />

Prostasinwaslocatedmainlyontheapicalplasmamembranealthoughminoramountscouldbedetectedonthebasolateralplasmamembrane(Fig.1,lanes9and10).These<br />

dataareconsistentwiththeexistingliteratureshowingthe<br />

Downloaded from www.jbc.org at DEF Consortium - Københavns Universitetsbibliotek, on June 15, 2011<br />

FEBRUARY 18, 2011 • VOLUME 286 • NUMBER 7 JOURNAL OF BIOLOGICAL CHEMISTRY 5795


Matriptase Activates Prostasin on the Basolateral Plasma Membrane<br />

FIGURE 1. Matriptase is located on the basolateral membrane whereas<br />

prostasin is apically located. Caco-2 cells grown on Transwell filters were<br />

biotin-labeled from either the apical (A) (lanes 1, 3, 5, 7, and 9) or basolateral<br />

(B) side (lanes 2, 4, 6, 8, and 10). Biotinylated proteins were precipitated with<br />

monomeric avidin, separated by SDS-PAGE and analyzed by Western blot<br />

with antibodies against total matriptase (M32) (lanes 1 and 2), matriptase-<br />

HAI-1 complex (M69) (lanes 3 and 4), the inhibitor HAI-1 (lanes 5– 8) and<br />

prostasin (lanes 9 and 10) / boiling of samples. Samples analyzed with<br />

the anti-prostasin antibody were boiled and reduced with DTT. The positions<br />

of molecular weight markers are indicated on the left. Protein and<br />

complex are marked with arrows and size on the right. The majority of the<br />

plasma membrane-bound matriptase was located on the basolateral membrane<br />

in a 70 kDa form, as detected with the antibody M32 (lane 2). A small<br />

fraction of the basolateral plasma membrane-bound matriptase was found<br />

in a 120 kDa form (lane 2). This form was also detected by the M69 (lane 4)<br />

and anti-HAI-1 (lane 6) antibodies. HAI-1 was found both on apical and basolateral<br />

plasma membrane (lanes 5– 8). Apical HAI-1 (lane 5) migrated<br />

faster on the SDS-PAGE than the basolateral HAI-1 (lane 6) when samples<br />

were not boiled but displayed the same size after boiling of the samples<br />

(lanes 7 and 8). Plasma membrane-bound prostasin was located mainly on<br />

the apical side (lane 9), although minor amounts were detected on the basolateral<br />

membrane (lane 10). Results shown are representative of five independent<br />

experiments.<br />

basolaterallocalizationofmatriptaseandapicallocalizationof<br />

prostasininapolarizedcell(15,18).<br />

Next,weinvestigatedthesubcellularsteadystatedistributionofmatriptaseandHAI-1inCaco-2cellsbyimmunocytochemistry(Fig.2).Caco-2cellsweregrownonTranswellfiltersbeforefixation,permeabilization,andimmunolabeling.<br />

MatriptaseandHAI-1werebothdetectedonthebasolateral<br />

plasmamembrane(Fig.2,A–F).Interestingly,matriptaseand<br />

HAI-1werealsodetectedinstructuresneartheapicalplasma<br />

membrane(Fig.2F).Onlyweakdetectionofmatriptase-<br />

HAI-1complexwasobservedonthebasolateralplasmamem-<br />

brane(Fig.2D).Surprisinglythemajorityofthematriptase-<br />

HAI-1complexwasdetectedinstructuresneartheapical<br />

plasmamembrane(Fig.2,D–F).Tofurtherinvestigatethe<br />

locationofthematriptase-HAI-1complexintheapicalregion<br />

ofthecells,Caco-2cellswereimmunolabeledwithandwithoutpermeabilizationoftheplasmamembrane(Fig.2,Gand<br />

H).Withoutpermeabilization,almostnomatriptase-HAI-1<br />

complexcouldbedetected(Fig.2H),however,withpermeabilizationadistinctlabelingofthevesicularstructureswas<br />

clearlyvisiblebelowtheapicalplasmamembrane(Fig.2G),<br />

confirminganintracellularlocalizationofthestructurescontainingthematriptase-HAI-1complex.Thesedataareconsistentwithourbiotinylationexperiments,aswewereonly<br />

abletolabelmatriptaseonthebasolateralplasmamembrane<br />

(Fig.1,lanes1and3).<br />

TheMatriptase-HAI-1ComplexIsGeneratedduring<br />

Endocytosis—Oursteadystateexperimentsshowthatalarge<br />

portionofmatriptaseispresentina70kDaformonthebasolateralplasmamembrane.Furthermore,theexperiments<br />

showthatlargeamountsofmatriptase-HAI-1complexare<br />

detectedinintracellularstructures.Thissuggeststhat<br />

matriptaseisendocytosedfromtheplasmamembraneand<br />

inhibitedbyHAI-1duringthisprocess.<br />

WeexaminedtheendocytosisofbasolateralplasmamembraneboundmatriptaseandHAI-1usingbiotinylationand<br />

internalizationtechniques(see“ExperimentalProcedures”).<br />

Theendocytosisexperimentsshowedthatmatriptaseisvery<br />

efficiently,andalmostcompletely,endocytosedwithin60min<br />

fromthebasolateralmembrane(Fig.3A).Furthermore,a<br />

3-foldincreaseinthesignalofmatriptase-HAI-1complexes<br />

wasdetectedwithinthefirst90minofincubation,showing<br />

thatthe70kDamatriptaseformsacomplexwithHAI-1duringtheendocytosis(Fig.3B).Severalmatriptase-HAI-1complexesbetween95and120kDaweregeneratedfromthe70<br />

kDaformduringtheendocytosis,asdetectedbyboththe<br />

matriptaseandHAI-1antibodies(Fig.3,A–C).After90min,<br />

allofthesurface-labeledmatriptasewasfoundincomplex<br />

withHAI-1andwaslocatedexclusivelyintracellularly<br />

(Fig.3B).<br />

The55kDaHAI-1hadanendocytosispatterndifferent<br />

fromHAI-1incomplexwithmatriptase.The55kDaHAI-1<br />

wasonlypartiallyendocytosed,withthelargestintracellular<br />

poolseenafter15min.ThissuggeststhatHAI-1,whennotin<br />

complexwithmatriptase,isrecyclingtotheplasmamembranefromearlyendosomes(Fig.3C).Thistypeofrecycling<br />

ofHAI-1hasbeenshownpreviouslyinMDCKcells(24).<br />

AHAI-1complexaround85kDawasobservedintheTotal<br />

panelfromtimepoint0to15min,whichwasnotdetectedby<br />

thematriptaseantibodies(Fig.3,compareCtoAandB,respectively).ThiscouldpossiblybeaHAI-1-prostasincomplex.Asimilarcomplexhaspreviouslybeenreportedinkeratinocytes(22).Togetherthesedatashowthatmatriptaseis<br />

efficientlyendocytosedfromthebasolateralsideandformsa<br />

complexwithHAI-1duringthisprocess.<br />

MatriptaseZymogenIsCleavedtotheActiveProteaseonthe<br />

SurfaceoftheBasolateralPlasmaMembrane—Wehaveupto<br />

nowshowedthatmatriptaseispresentpredominantlyina70<br />

kDaformonthebasolateralplasmamembraneandisefficientlyendocytosedformingacomplexwithHAI-1thataccumulatesinintracellularstructures,makingitaprerequisite<br />

thatmatriptaseactivationoccurspriortoendocytosis.<br />

Whethermatriptaseactivationoccursbeforearrivaloratthe<br />

plasmamembraneisnotwelldefined.<br />

Weexaminedtheactivationcleavageofbasolateralplasma<br />

membraneboundmatriptaseusingsurfacebiotinylationand<br />

anincubationassay.Underreducingconditions,matriptase<br />

zymogenisa70kDaproteinwhileenzymatically-active<br />

matriptaseseparatesintotwofragments;thestemdomain<br />

andthe30kDaproteasedomain.TheantibodyIM1014reacts<br />

withtheserineproteasedomainofbothzymogenandactivatedmatriptaseunderreducingconditions.Thisexperiment<br />

showedthatminoramountsofthe70kDamatriptasezymogenwerepresentonthebasolateralplasmamembranetogetherwith30kDacleavedmatriptaseserineproteasedomain<br />

(Fig.4,lane1).Thebasolateralmatriptasezymogenwasrapidlycleavedasonlythe30kDabandcouldbedetectedafter<br />

Downloaded from www.jbc.org at DEF Consortium - Københavns Universitetsbibliotek, on June 15, 2011<br />

5796 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 286 • NUMBER 7 • FEBRUARY 18, 2011


Matriptase Activates Prostasin on the Basolateral Plasma Membrane<br />

FIGURE 2. Matriptase-HAI-1 complex accumulates in intracellular structures. Caco-2 cells grown on Transwell filters, were fixed, permeabilized, and immunolabeled<br />

with antibodies against total matriptase (M32) and HAI-1 (A–C) or matriptase-HAI-1 complex (M69) and HAI-1 (D–F). Using a confocal scanning<br />

microscope, images were taken in the XY plane showing a single section through the monolayer and the XZ plane showing a cross section of the monolayer.<br />

The position of the plane of the XY section is indicated on the XZ plane (black arrows) on the right. The scale bar represents 10 m. A–C, Matriptase<br />

was detected on the basolateral plasma membrane of the Caco-2 cells co-localizing with HAI-1. Matriptase and HAI-1 was also co-localizing in structures<br />

near the apical membrane (white arrowheads). D–F, matriptase-HAI-1 complex was observed in structures near the apical plasma membrane (white arrowheads).<br />

Low amounts of matriptase-HAI-1 complex were detected on the basolateral plasma membrane. HAI-1 was detected both on apical and basolateral<br />

plasma membranes as well as in the apical structures co-localizing with matriptase. G and H, Caco-2 cells were treated with or without Triton X-100 prior to<br />

immunolabeling with the antibody M69. G, in the cells permeabilized with Triton X-100, a distinct detection of matriptase-HAI-1 complex was observed in<br />

apical vesicular structures. H, in cells without permeabilization, the vesicular structures with matriptase-HAI-1 complex were not detected. Only a<br />

weak basolateral signal was observed. The nuclei were visualized by DAPI staining shown in blue. Results shown are representative of three independent<br />

experiments.<br />

Downloaded from www.jbc.org at DEF Consortium - Københavns Universitetsbibliotek, on June 15, 2011<br />

just5minofincubation(Fig.4,lane2).Aslightincreasein<br />

the30kDabandintensitywasobservedupto15minafter<br />

labeling,mostlikelycausedbyahigheraffinityofIM1014<br />

antibodyforthecleaved30kDaproteasedomainthanthe70<br />

kDaform.Thesedatasuggestthatmatriptaseisfastandefficientlycleavedtoitsactiveformonthebasolateralcell<br />

surface.<br />

ActiveMatriptaseIsPresentontheBasolateralPlasma<br />

Membrane—Ithaspreviouslybeenreportedthattheperiodof<br />

timeforactivematriptasetoactonitssubstratesisverylimitedasmatriptaseactivationistightlycoupledtoinhibitionby<br />

HAI-1.Thiswindowofactionisassumedtobeinbetweenthe<br />

proteolyticcleavagecreatingthefullyactiveproteaseandthe<br />

rapidcomplexformationwithitsinhibitorHAI-1.Ourpreviousexperimentssuggestthatbothmatriptaseactivationand<br />

inhibitiontakesplaceonthebasolateralmembraneandwe<br />

wouldthereforeexpecttofindfreeactivematriptaseonthe<br />

basolateralplasmamembrane.Toaddressthis,weinvestigatedwhethermatriptaseatthebasolateralplasmamembrane<br />

wasabletobindtothegeneralserineproteaseinhibitors,<br />

aprotininandleupeptin,towhichweexpectonlytheactive<br />

proteasetobind.Wewereabletopurifymatriptasefromthe<br />

FEBRUARY 18, 2011 • VOLUME 286 • NUMBER 7 JOURNAL OF BIOLOGICAL CHEMISTRY 5797


Matriptase Activates Prostasin on the Basolateral Plasma Membrane<br />

FIGURE 3. Matriptase in complex with HAI-1 is generated during the efficient endocytosis of matriptase from the basolateral plasma membrane.<br />

Caco-2 cells were grown on Transwell filters, and proteins on the basolateral plasma membrane were biotinylated at 4 °C using a cleavable biotinylation<br />

reagent, s-NHS-SS-biotin. The labeled cells were incubated at 37 °C for the time indicated (0 –120 min). After incubation, the proteins remaining on the<br />

plasma membrane were biotin-stripped, using the membrane non-permeant reducing agent glutathione, leaving only endocytosed proteins biotinylated<br />

(endocytosis panels). Biotin reduction was omitted in a parallel set of samples to monitor the degradation of the biotinylated proteins over time (total panels).<br />

Biotinylated proteins were precipitated with monomeric avidin agarose, and the avidin pull-downs were analyzed with SDS-PAGE and Western blotting<br />

using the antibodies (A) M32 against total matriptase, (B) M69 against matriptase-HAI-1 complex, and (C) M19 against HAI-1. The positions of molecular<br />

markers (kDa) are indicated on the left. Proteins and complexes are indicated with arrows and size. Quantification of bands from Western blots was done<br />

using the software ImageJ. A graphic presentation of three independent endocytosis experiments is shown on the right hand side. The dotted line equals<br />

total protein and solid line equals endocytosed protein. The standard deviation is shown with error bars. A, matriptase was endocytosed from the basolateral<br />

plasma membrane within 60 min. Approximately 80% of matriptase was endocytosed at 60 min as indicated on the graph. B, during the 120-min incubation<br />

there was a 3-fold increase in the matriptase-HAI-1 complexs. C, free 55 kDa HAI-1 was partially endocytosed within 15 min. (C, endocytosis panel).<br />

Results shown are representative of three independent experiments.<br />

basolateralplasmamembrane,withbothaprotinin-andleupeptin-coupledSepharose(Fig.5A,lanes3and5).No<br />

matriptasewaspurifiedwiththenegativecontrolsoybean<br />

trypsininhibitor-SepharoseoruncoupledSepharose(datanot<br />

shown).Theinhibitor-boundfractionofmatriptasewasnot<br />

detectedbytheantibodydetectingmatriptase-HAI-1complexes(Fig.5B,lanes3and5).<br />

Wewantedtoverifythatthepurifiedmatriptasedetected<br />

usingtheinhibitor-coupledSepharoseindeedwasfreeactive<br />

matriptaseandnotactivatedmatriptasedissociatedfromthe<br />

matriptase-HAI-1complexes.Totestthis,asamplecontainingmostlymatriptase-HAI-1complexwasexposedtothe<br />

inhibitor-coupledSepharose.Nodetectablematriptasefrom<br />

thematriptase-HAI-1complexwaspurifiedwiththeinhibitor<br />

Sepharose(Fig.5,lanes4and6).Thisverifiesthatitisfree<br />

activematriptasethatisbeingpulleddownandthatthe<br />

matriptase-HAI-1complexisnotdissociatedduringourcell<br />

extractionprocedure.<br />

Finally,totesttheproteolyticactivityofmatriptase,thetwo<br />

biotinylatedfractions,containing70kDamatriptaseand<br />

matriptase-HAI-1complexes,respectively,wereanalyzedfor<br />

theirabilitytodisplaygelatinolyticactivityatpH8.75,where<br />

matriptasehasoptimalenzymaticactivity(25)(Fig.5C).The<br />

fractionwiththe70kDamatriptasedisplayedonegelatinolyticbandaround70kDaatpH8.75,correspondingtothe<br />

sizeoffreeactivematriptase(Fig.5C,lane1).Thefraction<br />

containingmostlymatriptase-HAI-1complex,showedonly<br />

weakgelatinolyticactivityaround70kDa(Fig.5C,lane2).<br />

Thus,thegelatinolyticactivitypatternmatchestheoneobservedformatriptaseimmunoreactivitywiththeantibody<br />

M32(Fig.5A)implyingthatthegelatinolyticactivityobserved<br />

iscausedbymatriptase.<br />

Downloaded from www.jbc.org at DEF Consortium - Københavns Universitetsbibliotek, on June 15, 2011<br />

5798 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 286 • NUMBER 7 • FEBRUARY 18, 2011


Matriptase Activates Prostasin on the Basolateral Plasma Membrane<br />

FIGURE 4. Matriptase is cleaved to the two chain form on the basolateral<br />

cell surface. Caco-2 cells were grown on Transwell filters, biotinylated<br />

from the basolateral side at 4 °C and incubated up to 60 min at 37 °C after<br />

labeling. Biotinylated proteins were pulled down with streptavidin-agarose,<br />

boiled, reduced and analyzed by SDS-PAGE and Western blotting using the<br />

antibody IM1014. The positions of molecular markers (kDa) are indicated on<br />

the left. Bands are indicated with arrows and size. At time 0 the matriptase is<br />

detected as a 70 kDa band, representing the non-cleaved zymogen and a<br />

30 kDa band representing the cleaved protease domain (lane 1). After 5 min<br />

of incubation the zymogen could no longer be detected and an increase in<br />

the 30 kDa serine protease domain was detected (lanes 2– 6). Results shown<br />

are representative of two independent experiments.<br />

FIGURE 5. Matriptase is active on the basolateral plasma membrane<br />

but not intracellularly. Caco-2 cells on Transwell filters were surfacebiotinylated<br />

from the basolateral side at 4 °C. Some were lysed immediately<br />

after biotinylation (0) and some were incubated for 2 h at 37 °C to<br />

transform all biotin-labeled matriptase into activated matriptase in complex<br />

with HAI-1 (2). Biotinylated proteins were precipitated with monomeric<br />

avidin and gently eluted with biotin. The avidin pull-downs were<br />

divided into three groups: No further treatment (lanes 1 and 2), pulldown<br />

with aprotinin-Sepharose (lanes 3 and 4) and pull-down with leupeptin-Sepharose<br />

(lanes 5 and 6). The samples were analyzed by SDS-<br />

PAGE and Western blot with antibodies against total matriptase (M32)<br />

and matriptase-HAI-1 complex (M69). A, only at time 0 was it possible to<br />

pull-down M32-detectable matriptase with both aprotinin and leupeptin.<br />

B, M69-detectable matriptase was pulled down with the monomeric<br />

avidin (B, lanes 1 and 2), but was lost with additional inhibitor pull-down<br />

(B, lanes 3– 6). C, two avidin-purified fractions showed gelatinolytic properties<br />

with a band around 70 kDa (C, lanes 1 and 2), matching the size<br />

and pattern of A, lanes 1 and 2. Results shown are representative of two<br />

independent experiments.<br />

FIGURE 6. Active prostasin is present on the basolateral as well as on<br />

the apical plasma membrane. Caco-2 cells on Transwell filters were surface<br />

biotinylated from either the apical or basolateral side at 4 °C. Biotinylated<br />

proteins were precipitated with monomeric avidin and gently eluted<br />

with biotin. The avidin pull-downs were divided into three: No further treatment<br />

(lanes 1 and 2), pull-down with aprotinin-Sepharose (lanes 3 and 4)<br />

and pull-down with trypsin inhibitor as negative control (lanes 5 and 6). The<br />

pull-down fractions were analyzed by SDS-PAGE and Western blotting.<br />

Prostasin was purified from both apical plasma membrane (lane 1) and basolateral<br />

plasma membrane (lane 2). The majority of the apical prostasin<br />

could be pulled down with the serine protease inhibitor aprotinin (lane 3). A<br />

small but significant fraction of basolateral prostasin was also able to bind<br />

aprotinin and hence was in its active form (lane 4). No prostasin binding was<br />

observed when using the negative control trypsin-coupled Sepharose<br />

(lanes 5 and 6). Results shown are representative of three independent<br />

experiments.<br />

ActiveProstasinIsPresentonBoththeApicalandtheBasolateralPlasmaMembrane—Allofourdatasuggestthat<br />

matriptaseisactiveandabletocleaverelevantsubstratesina<br />

shortperiodoftimeonthebasolateralplasmamembranebeforeitformsacomplexwithHAI-1andisendocytosed.If<br />

prostasinisindeedasubstrateformatriptaseinpolarizedepithelialcells,wewouldexpecttofindthecleavedformofprostasinonthebasolateralmembrane.Totestthis,weutilized<br />

thefactthatprostasinisaserineproteaseandonlythe<br />

cleavedformisproteolyticallyactiveandtherebyabletobind<br />

serineproteaseinhibitorssuchasaprotinin.Ourexperiment<br />

showedthatprostasinonboththeapicalandthebasolateral<br />

plasmamembranewasabletobindtoaprotinin(Fig.6,lanes<br />

3and4).Thissuggeststhatprostasinisactiveonboththe<br />

apicalandthebasolateralsideofthecell.<br />

MatriptaseCo-localizeswithProstasinontheBasolateral<br />

PlasmaMembrane,SubsequentlyProstasinIsTranscytosedto<br />

theApicalPlasmaMembrane—Weknowfromthesurface<br />

labelingexperimentthatthemajorityofmembrane-bound<br />

prostasinatsteadystateislocatedontheapicalside(Fig.1,<br />

lane9)whereasonlyaminorfractionislocatedonthebasolateralmembraneinCaco-2cells(Fig.1,lane10).Fromour<br />

experiments,wealsoknowthatprostasinisnotonlyfoundin<br />

itsactiveformontheapicalplasmamembranebutalsoonthe<br />

basolateralplasmamembrane.Thiscouldindicatethatprostasinistransportedtothebasolateralplasmamembraneto<br />

getactivatedbymatriptase.<br />

Wewantedtoinvestigatetheendocytictransportofthe<br />

apicalandbasolateralprostasin.Initially,wetestedifthe<br />

twomembranefractionswereendocytosedusingthesame<br />

experimentalsetupasformatriptaseandHAI-1endocytosis.<br />

Wefoundthatprostasinwasnotendocytosedfromtheapical<br />

plasmamembranebutwascompletelyendocytosedfromthe<br />

basolateralplasmamembranewithin60min(datanot<br />

shown).Thismadeusquestionifprostasinisinitiallytransportedtothebasolateralmembraneandactivatedby<br />

matriptasebeforeitisre-routedbytranscytosistotheapical<br />

membrane.Alongresidencetimeattheapicalplasmamem-<br />

Downloaded from www.jbc.org at DEF Consortium - Københavns Universitetsbibliotek, on June 15, 2011<br />

FEBRUARY 18, 2011 • VOLUME 286 • NUMBER 7 JOURNAL OF BIOLOGICAL CHEMISTRY 5799


Matriptase Activates Prostasin on the Basolateral Plasma Membrane<br />

FIGURE 7. Prostasin and HAI-1 are transcytosed from the basolateral to the apical plasma membrane. Caco-2 cells were surface biotinylated from the<br />

basolateral side at 4 °C. The labeled cells were incubated at 37 °C for the time indicated in the internalization panel (4, 12, or 18 h as indicated). After incubation<br />

the proteins remaining on the plasma membrane were biotin-stripped, using the membrane non-permeant reducing agent glutathione, from either<br />

apical (A red.), basolateral (B red.), or both sides (A B red.). Biotin reduction was omitted in a set of samples to monitor the total amount of biotinylated<br />

protein remaining after the incubation (Total). The experiment was performed in duplicates. Biotinylated proteins were precipitated with monomeric avidin<br />

agarose, separated with SDS-PAGE and analyzed by Western blotting. A, prostasin could be biotin-stripped from the apical side (A red.) after 18 h of internalization,<br />

as a decrease in signal was observed compared with the total samples. This shows that prostasin has moved from the basolateral to the apical side.<br />

When surface proteins were biotin-stripped from the basolateral side no decrease in signal was observed (B red.), once again showing that no biotinylated<br />

prostasin was left on the basolateral side after 18 h. To clarify if some of the prostasin was located intracellularly, the cells were biotin stripped from both<br />

apical and basolateral side (A B red.). Biotin-stripping from both apical and basolateral side removed the majority of the biotinylated prostasin suggesting<br />

that all of prostasin had within the 18 h transferred from the basolateral to the apical plasma membrane. B, HAI-1 was transcytosed from the basolateral to<br />

the apical plasma membrane within 12 h. Most of HAI-1 could be biotin stripped from the apical side after incubation (A red.) where a major fraction was left<br />

after basolateral reduction (B. red) showing that a large fraction of biotinylated HAI-1 had moved from the basolateral to the apical plasma membrane. It<br />

was demonstrated that the faster migrating form of HAI-1 resides at the apical plasma membrane 12 h after biotinylation and the slower migrating form<br />

remained at the basolateral plasma membrane, as these could be biotin-stripped from their respective sites (A red. compared with B red., respectively).<br />

C, after just 4 h of incubation most biotinylated matriptase was no longer detectable in the cell. The remaining matriptase was mainly found in the 120 kDa<br />

complex with HAI-1. A fraction of the matriptase-HAI-1 complex could be biotin stripped from the basolateral side (B red.) but not the apical side (A red.).<br />

D, M69 detectable matriptase showed the same amount of matriptase-HAI-1 complex at time 0 as for 4 h of incubation. A fraction of the complex was reducible<br />

from the basolateral side after 4 h of incubation (B red.). Results shown are representative of three independent experiments.<br />

branecombinedwithashortresidencetimeatthebasolateral<br />

plasmamembranewouldgiveasteadystateaccumulationat<br />

theapicalplasmamembrane.We,therefore,testedifthebasolaterallyendocytosedfractionwastranscytosedtotheapical<br />

membrane.<br />

Transcytosisforseveralincubationtimeswasinvestigated<br />

(1,2,4,8,12,and18h)ThetimepointsmostclearlydemonstratingtranscytosisareshowninFig.7.Transcytosisofprostasinwasdetectableafter8h(datanotshown)andmostof<br />

prostasinwastranscytosedfromthebasolateraltotheapical<br />

plasmamembraneafter18h(Fig.7A).HAI-1wasalsotranscytosed,however,lessefficientlythanprostasin(Fig.7B).<br />

Ourresultsstronglysuggestthattheslowmigratingformof<br />

HAI-1atthebasolateralplasmamembraneistheprecursor<br />

forthefastermigratingHAI-1presentontheapicalplasma<br />

membrane(Fig.7B).<br />

Matriptase-HAI-1 complex has been purified from milk,<br />

suggesting that matriptase is secreted from the apical<br />

plasma membrane (16). However, we were unable to detect<br />

transcytosis of matriptase from the basolateral to the apical<br />

plasma membrane, as matriptase has a shorter half-life in<br />

the cell than HAI-1 and prostasin (Fig. 7, C and D). The<br />

remaining biotin-labeled matriptase was found in the<br />

matriptase-HAI-1 complexes between 95 and 120 kDa after<br />

4 h incubation (Fig. 7, C and D). These complexes were<br />

also detected with the anti-HAI-1 antibody after 4 h incubation<br />

(data not shown). Instead, basolaterally biotin-labeled<br />

matriptase could be detected in the basolateral media<br />

Downloaded from www.jbc.org at DEF Consortium - Københavns Universitetsbibliotek, on June 15, 2011<br />

5800 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 286 • NUMBER 7 • FEBRUARY 18, 2011


Matriptase Activates Prostasin on the Basolateral Plasma Membrane<br />

A<br />

B<br />

C<br />

D<br />

Total matriptase<br />

(M32)<br />

Matriptase-HAI-1<br />

complex (M69)<br />

HAI-1<br />

Prostasin<br />

5m 15m 30m 1h 2h 4h 8h 18h<br />

- 120 kD<br />

Apical media<br />

- 70 kD<br />

- 120 kD<br />

- 70 kD<br />

- 120 kD<br />

- 50 kD<br />

- 40 kD<br />

FIGURE 8. Matriptase is shed into the basolateral media while basolateral<br />

prostasin is transcytosed and shed into the apical media. Caco-2<br />

cells were surface-biotinylated from the basolateral side at 4 °C. Cells were<br />

incubated at 37 °C after labeling and apical and basolateral media was collected<br />

at 5, 15, 30 min, 1, 2, 4, 8, and 18 h to reveal any shedding of the biotin-labeled<br />

matriptase, HAI-1 and prostasin. Biotinylated proteins from the<br />

media were pulled down with monomeric avidin agarose, separated with<br />

SDS-PAGE and analyzed by Western blotting. A, matriptase was released to<br />

the basolateral media within 1 h as three forms: a 70 kDa form and two proteolytically<br />

shed complexes at 85 and 110 kDa. B, two complexes of size 85<br />

and 110 kDa could be detected with the M69 antibody within 1–2 h, suggesting<br />

these to be matriptase-HAI-1 complexes. C, no free 55 kDa HAI-1<br />

was found in the media, only HAI-1 in the two complexes at 85 and 110 kDa<br />

was detected in the basolateral media after 1 h and accumulated up to 18 h.<br />

D, basolateral prostasin was detected in the apical media after 18 h incubation,<br />

confirming transcytosis before secretion into the apical media. Results<br />

shown are representative of two independent experiments.<br />

after only 1 h and matriptase accumulated in the media<br />

over the 18 h (Fig. 8, A and B). Both the 70 kDa form as<br />

well as two forms of the matriptase-HAI-1 complex around<br />

110 and 85 kDa were detected in the media. Only HAI-1 in<br />

complex with matriptase was detected in the basolateral<br />

media (Fig. 8C), whereas no free HAI-1 could be detected.<br />

None of the biotin-labeled matriptase or HAI-1 could be<br />

detected in the apical media (data not shown). Basolateral<br />

prostasin could not be detected in the basolateral media<br />

(data not shown) but after 18 h of incubation after the biotinylation,<br />

prostasin appeared in the apical media (Fig. 8D).<br />

This is consistent with the finding that prostasin is transcytosed<br />

from the basolateral to the apical plasma membrane<br />

within this time frame. Thus, we were able to show that<br />

prostasin on the basolateral plasma membrane is transcytosed<br />

to the apical plasma membrane from where it is shed<br />

to the media. This suggests that prostasin is routed via the<br />

basolateral plasma membrane where it is activated before it<br />

is transcytosed to the apical membrane, thereby providing<br />

a mechanism for activation of the prostasin zymogen in<br />

polarized epithelial cells by matriptase or other basolaterally<br />

located serine proteases, despite the separate steadystate<br />

localization of the proteases.<br />

DISCUSSION<br />

Ithasbeenshownthatmatriptaseandprostasinareconstitutivelyexpressedandco-localizeinmostepitheliaincluding<br />

thetissuesaffectedbymatriptaseablation,suggestingapossibleglobalroleforamatriptase-prostasincascadeinepithelial<br />

homeostasis(26).Inapolarizedcell,matriptaseisatsteadystateconcentratedandlocalizedmainlyatthebasolateral<br />

plasmamembranetogetherwithHAI-1(15,24)whileprostasinismainlyconcentratedandlocalizedtotheapicalplasma<br />

membrane(17,18).Ourdataprovideapossiblesolutionto<br />

howmatriptasecanactivateprostasindespitetheirtwodifferentsubcellularlocalizations,byshowingthatthetwomoleculesmeetenroute.<br />

Matriptaseismainlylocatedonthebasolateralplasma<br />

membranewhereitisactivated,inhibited,endocytosedin<br />

complexwithitsinhibitorHAI-1andisaccumulatedinintracellularstructures.ObservationofthetransmembraneNterminalfragmentofmatriptaseinintracellularcompartmentshaspreviouslybeendescribedinratenterocytes(14).<br />

Weshowherethatprostasinismainlylocatedontheapical<br />

plasmamembraneinitsactiveform.Interestingly,wefinda<br />

smallfractionofprostasinonthebasolateralplasmamembraneco-localizingwithmatriptase,makingitpossiblefor<br />

proteaseandsubstratetointeractandactivationtooccur.In<br />

agreementwiththis,weshowthatpartofthebasolateral<br />

prostasinisactivebyitsabilitytobindtogeneralserineproteaseinhibitors.Thebasolateralprostasinisendocytosedand<br />

transcytosedtotheapicalplasmamembrane,whereit<br />

accumulates.<br />

Bothmatriptaseandprostasinhasbeenshowntobeableto<br />

activatetheepithelialsodiumchannel(ENaC)inXenopus<br />

oocytes(5,27,28).ENacisanepithelialmembrane-bound<br />

sodiumchannellocatedintheapicalmembraneofpolarized<br />

cellsandisrequiredfornormalepidermaldifferentiation(29,<br />

30).Becausematriptaseistargetedtothebasolateralmembranewhereitisactivatedandrapidlyinhibiteditismore<br />

likelythatprostasinwhichco-localizeswithENaContheapicalmembraneisacandidateactivatorinpolarizedcells.Both<br />

matriptaseandprostasinhavebeencoupledtomaintenance<br />

offunctionaltightjunctionsinavarietyofepithelialcelltypes.<br />

InaCaco-2model,lossofmatriptasewasassociatedwithenhancedexpressionandincorporationofthepore-forming<br />

proteinclaudin-2attightjunctions(2).Prostasinhasbeen<br />

reportedtoregulatetightjunctions,paracellularpermeability<br />

andTEERbyaproteaseactivity-dependentmechanismin<br />

renalcells(4,5).Thismeansthatthetwoproteasesareboth<br />

involvedincascadesimportantfortightjunctionformationin<br />

polarizedepithelia.<br />

Increasingevidencesuggeststhatserineproteasescontributeinacomplexwaytotheregulationofintestinalintegrity<br />

andbarrierfunction.Proteaseinhibitorshavebeenshownto<br />

suppresstheformationoftightjunctionsingastrointestinal<br />

celllinessuggestingthatproteolyticactivityisnecessaryfora<br />

functionalepithelialbarrier(31).Theepithelialbarrieriscrucialinthegastrointestinaltractandisoftencompromisedin<br />

inflammatoryboweldiseaseslikeCrohndiseaseandulcerativecolitis.ThesodiumchannelENaCisdown-regulatedin<br />

Downloaded from www.jbc.org at DEF Consortium - Københavns Universitetsbibliotek, on June 15, 2011<br />

FEBRUARY 18, 2011 • VOLUME 286 • NUMBER 7 JOURNAL OF BIOLOGICAL CHEMISTRY 5801


Matriptase Activates Prostasin on the Basolateral Plasma Membrane<br />

inflammatoryboweldiseasesandhasbeenproposedasatherapeutictarget(32–34).Itwillbeimportantforfuturestudies<br />

toidentifytheroleofthematriptase-prostasincascadeand<br />

thesubstratesinvolvedinthispathwaytoinvestigatetherole<br />

ofmatriptaseininflammatorydiseasesofthegastrointestinal<br />

tract.<br />

Together,ourdatashowthatactivematriptaseanditssubstrateprostasinco-localizeatthebasolateralplasmamembraneandherebyprovideavenueforamatriptase-prostasin<br />

zymogencascadetobeinitiatedinpolarizedepithelialcells.<br />

Furthermore,weshowhowprostasinistransportedafteractivationtotheapicalmembranetoco-localizewithitsprimary<br />

substrateENaC.<br />

Acknowledgment—WethankDr.MaryJoDantonforcritically<br />

readingthemanuscript.<br />

REFERENCES<br />

1. List,K.,Kosa,P.,Szabo,R.,Bey,A.L.,Wang,C.B.,Molinolo,A.,and<br />

Bugge,T.H.(2009)Am.J.Pathol.175,1453–1463<br />

2. Buzza,M.S.,Netzel-Arnett,S.,Shea-Donohue,T.,Zhao,A.,Lin,C.Y.,<br />

List,K.,Szabo,R.,Fasano,A.,Bugge,T.H.,andAntalis,T.M.(2010)<br />

Proc.Natl.Acad.Sci.U.S.A.107,4200–4205<br />

3. Yu,J.X.,Chao,L.,andChao,J.(1995)J.Biol.Chem.270,13483–13489<br />

4. Steensgaard,M.,Svenningsen,P.,Tinning,A.R.,Nielsen,T.D.,Jorgensen,F.,Kjaersgaard,G.,Madsen,K.,andJensen,B.L.(2010)Acta<br />

Physiol.200,347–349<br />

5. Verghese,G.M.,Gutknecht,M.F.,andCaughey,G.H.(2006)Am.J.<br />

Physiol.CellPhysiol.291,C1258–C1270<br />

6. Leyvraz,C.,Charles,R.P.,Rubera,I.,Guitard,M.,Rotman,S.,Breiden,<br />

B.,Sandhoff,K.,andHummler,E.(2005)J.CellBiol.170,487–496<br />

7. List,K.,Szabo,R.,Wertz,P.W.,Segre,J.,Haudenschild,C.C.,Kim,<br />

S.Y.,andBugge,T.H.(2003)J.CellBiol.163,901–910<br />

8. Ovaere,P.,Lippens,S.,Vandenabeele,P.,andDeclercq,W.(2009)<br />

TrendsBiochem.Sci.34,453–463<br />

9. Szabo,R.,Kosa,P.,List,K.,andBugge,T.H.(2009)Am.J.Pathol.179,<br />

2015–2022<br />

10. List,K.,Haudenschild,C.C.,Szabo,R.,Chen,W.,Wahl,S.M.,Swaim,<br />

W.,Engelholm,L.H.,Behrendt,N.,andBugge,T.H.(2002)Oncogene<br />

21,3765–3779<br />

11. Netzel-Arnett,S.,Currie,B.M.,Szabo,R.,Lin,C.Y.,Chen,L.M.,Chai,<br />

K.X.,Antalis,T.M.,Bugge,T.H.,andList,K.(2006)J.Biol.Chem.281,<br />

32941–32945<br />

12. Alef,T.,Torres,S.,Hausser,I.,Metze,D.,Türsen,U.,Lestringant,G.G.,<br />

andHennies,H.C.(2009)J.Invest.Dermatol.129,862–869<br />

13. LeBivic,A.,Quaroni,A.,Nichols,B.,andRodriguez-Boulan,E.(1990)<br />

J.CellBiol.111,1351–1361<br />

14. Tsuzuki,S.,Murai,N.,Miyake,Y.,Inouye,K.,Hirayasu,H.,Iwanaga,T.,<br />

andFushiki,T.(2005)Biochem.J.388,679–687<br />

15. Wang,J.K.,Lee,M.S.,Tseng,I.C.,Chou,F.P.,Chen,Y.W.,Fulton,A.,<br />

Lee,H.S.,Chen,C.J.,Johnson,M.D.,andLin,C.Y.(2009)Am.J.<br />

Physiol.CellPhysiol.279,C459–470<br />

16. Lin,C.Y.,Anders,J.,Johnson,M.,andDickson,R.B.(1999)J.Biol.<br />

Chem.274,18237–18242<br />

17. Chen,M.,Chen,L.M.,Lin,C.Y.,andChai,K.X.(2008)Biochim.Biophys.Acta1785,896–903<br />

18. Selzer-Plon,J.,Bornholdt,J.,Friis,S.,Bisgaard,H.C.,Lothe,I.M.,Tveit,<br />

K.M.,Kure,E.H.,Vogel,U.,andVogel,L.K.(2009)BMC.Cancer9,<br />

201<br />

19. Lin,C.Y.,Wang,J.K.,Torri,J.,Dou,L.,Sang,Q.A.,andDickson,R.B.<br />

(1997)J.Biol.Chem.272,9147–9152<br />

20. Satomi,S.,Yamasaki,Y.,Tsuzuki,S.,Hitomi,Y.,Iwanaga,T.,and<br />

Fushiki,T.(2001)Biochem.Biophys.Res.Commun.287,995–1002<br />

21. Tseng,I.C.,Chou,F.P.,Su,S.F.,Oberst,M.,Madayiputhiya,N.,Lee,<br />

M.S.,Wang,J.K.,Sloane,D.E.,Johnson,M.,andLin,C.Y.(2008)<br />

Am.J.Physiol.CellPhysiol.295,C423–C431<br />

22. Chen,Y.W.,Wang,J.K.,Chou,F.P.,Chen,C.Y.,Rorke,E.A.,Chen,<br />

L.M.,Chai,K.X.,Eckert,R.L.,Johnson,M.D.,andLin,C.Y.(2010)<br />

J.Biol.Chem.285,31755–31762<br />

23. Lin,C.Y.,Tseng,I.C.,Chou,F.P.,Su,S.F.,Chen,Y.W.,Johnson,<br />

M.D.,andDickson,R.B.(2008)FrontBiosci.13,621–635<br />

24. Godiksen,S.,Selzer-Plon,J.,Pedersen,E.D.,Abell,K.,Rasmussen,H.B.,<br />

Szabo,R.,Bugge,T.H.,andVogel,L.K.(2008)Biochem.J.413,251–259<br />

25. Beliveau,F.,Desilets,A.,andLeduc,R.(2009)FEBSJ.276,2213–2226<br />

26. List,K.,Hobson,J.P.,Molinolo,A.,andBugge,T.H.(2007)J.Cell.<br />

Physiol.213,237–245<br />

27. Vallet,V.,Chraibi,A.,Gaeggeler,H.P.,Horisberger,J.D.,andRossier,<br />

B.C.(1997)Nature389,607–610<br />

28. Andreasen,D.,Vuagniaux,G.,Fowler-Jaeger,N.,Hummler,E.,and<br />

Rossier,B.C.(2006)J.Am.Soc.Nephrol.17,968–976<br />

29. Myerburg,M.M.,Harvey,P.R.,Heidrich,E.M.,Pilewski,J.M.,andButterworth,M.(2010)Am.J.Respir.CellMol.Biol.43,712–719<br />

30. Mauro,T.,Guitard,M.,Behne,M.,Oda,Y.,Crumrine,D.,Komuves,L.,<br />

Rassner,U.,Elias,P.M.,andHummler,E.(2002)J.Invest.Dermatol.<br />

118,589–594<br />

31. Bacher,A.,Griebl,K.,Mackamul,S.,Mitreiter,R.,Mückter,H.,and<br />

Ben-Shaul,Y.(1992)Exp.CellRes.200,97–104<br />

32. Farkas,K.,Yeruva,S.,Rakonczay,Z.,Jr.,Ludolph,L.,Molnar,T.,Nagy,<br />

F.,Szepes,Z.,Schnur,A.,Wittmann,T.,Hubricht,J.,Riederer,B.,Venglovecz,V.,Lazar,G.,Kiraly,M.,Zsembery,A.,Varga,G.,Seidler,U.,<br />

andHegyi,P.(2010)Inflamm.Bowel.Dis.inpress<br />

33. Sullivan,S.,Alex,P.,Dassopoulos,T.,Zachos,N.C.,Iacobuzio-<br />

Donahue,C.,Donowitz,M.,Brant,S.R.,Cuffari,C.,Harris,M.L.,Datta,<br />

L.W.,Conklin,L.,Chen,Y.,andLi,X.(2009)Inflamm.Bowel.Dis.15,<br />

261–274<br />

34. Zeissig,S.,Bergann,T.,Fromm,A.,Bojarski,C.,Heller,F.,Guenther,U.,<br />

Zeitz,M.,Fromm,M.,andSchulzke,J.D.(2008)Gastroenterology134,<br />

1436–1447<br />

Downloaded from www.jbc.org at DEF Consortium - Københavns Universitetsbibliotek, on June 15, 2011<br />

5802 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 286 • NUMBER 7 • FEBRUARY 18, 2011


Paper II<br />

Reduced prostasin (CAP1/PRSS8) activity eliminates HAI-1 and HAI-2 deficiency-associated<br />

developmental defects by preventing matriptase activation<br />

Roman Szabo 1 , Katiuchia Uzzun Sales 1 , Peter Kosa 1 , Natalia A. Shylo 1 , Sine Godiksen 1,2,3 ,<br />

Karina K. Hansen 1 , Stine Friis 1 , J. Silvio Gutkind 1 , Lotte K. Vogel 2 , Edith Hummler 4 , Eric<br />

Camerer 5,6 , and Thomas H. Bugge 1<br />

1 Oral and Pharyngeal Cancer Branch, National <strong>Institut</strong>e of Dental and Craniofacial Research,<br />

National <strong>Institut</strong>es of Health, Bethesda, MD, USA.<br />

2 Department of Cellular and Molecular Medicine, University of Copenhagen, Copenhagen,<br />

Denmark.<br />

3 Department of Biology, University of Copenhagen, Copenhagen, Denmark.<br />

4 Pharmacology and Toxicology Department, University de Lausanne, Lausanne, Switzerland.<br />

5 INSERM U970, Paris Cardiovascular Research Centre, Paris, F-75015, France. 6Université Paris-<br />

Descartes, Paris, F-75006, France.<br />

Published in PLoS Genetic, August 2012<br />

47


Reduced Prostasin (CAP1/PRSS8) Activity Eliminates HAI-<br />

1 and HAI-2 Deficiency–Associated Developmental<br />

Defects by Preventing Matriptase Activation<br />

Roman Szabo 1 , Katiuchia Uzzun Sales 1 , Peter Kosa 1 , Natalia A. Shylo 1 , Sine Godiksen 1,2,3 ,<br />

Karina K. Hansen 1 , Stine Friis 1 , J. Silvio Gutkind 1 , Lotte K. Vogel 2 , Edith Hummler 4 , Eric Camerer 5,6 ,<br />

Thomas H. Bugge 1 *<br />

1 Oral and Pharyngeal Cancer Branch, National <strong>Institut</strong>e of Dental and Craniofacial Research, National <strong>Institut</strong>es of Health, Bethesda, Maryland, United States of America,<br />

2 Department of Cellular and Molecular Medicine, Faculty of Health Science, University of Copenhagen, Copenhagen, Denmark, 3 Department of Biology, Faculty of<br />

Science, University of Copenhagen, Copenhagen, Denmark, 4 Pharmacology and Toxicology Department, University de Lausanne, Lausanne, Switzerland, 5 INSERM U970,<br />

Paris Cardiovascular Research Centre, Paris, France, 6 Université Paris-Descartes, Paris, France<br />

Abstract<br />

Loss of either hepatocyte growth factor activator inhibitor (HAI)-1 or -2 is associated with embryonic lethality in mice, which<br />

can be rescued by the simultaneous inactivation of the membrane-anchored serine protease, matriptase, thereby<br />

demonstrating that a matriptase-dependent proteolytic pathway is a critical developmental target for both protease<br />

inhibitors. Here, we performed a genetic epistasis analysis to identify additional components of this pathway by generating<br />

mice with combined deficiency in either HAI-1 or HAI-2, along with genes encoding developmentally co-expressed<br />

candidate matriptase targets, and screening for the rescue of embryonic development. Hypomorphic mutations in Prss8,<br />

encoding the GPI-anchored serine protease, prostasin (CAP1, PRSS8), restored placentation and normal development of<br />

HAI-1–deficient embryos and prevented early embryonic lethality, mid-gestation lethality due to placental labyrinth failure,<br />

and neural tube defects in HAI-2–deficient embryos. Inactivation of genes encoding c-Met, protease-activated receptor-2<br />

(PAR-2), or the epithelial sodium channel (ENaC) alpha subunit all failed to rescue embryonic lethality, suggesting that<br />

deregulated matriptase-prostasin activity causes developmental failure independent of aberrant c-Met and PAR-2 signaling<br />

or impaired epithelial sodium transport. Furthermore, phenotypic analysis of PAR-1 and matriptase double-deficient<br />

embryos suggests that the protease may not be critical for focal proteolytic activation of PAR-2 during neural tube closure.<br />

Paradoxically, although matriptase auto-activates and is a well-established upstream epidermal activator of prostasin,<br />

biochemical analysis of matriptase- and prostasin-deficient placental tissues revealed a requirement of prostasin for<br />

conversion of the matriptase zymogen to active matriptase, whereas prostasin zymogen activation was matriptaseindependent.<br />

Citation: Szabo R, Uzzun Sales K, Kosa P, Shylo NA, Godiksen S, et al. (2012) Reduced Prostasin (CAP1/PRSS8) Activity Eliminates HAI-1 and HAI-2 Deficiency–<br />

Associated Developmental Defects by Preventing Matriptase Activation. PLoS Genet 8(8): e1002937. doi:10.1371/journal.pgen.1002937<br />

Editor: Hamish S. Scott, SA Pathology, Australia<br />

Received April 5, 2012; Accepted July 18, 2012; Published August 30, 2012<br />

This is an open-access article, free of all copyright, and may be freely reproduced, distributed, transmitted, modified, built upon, or otherwise used by anyone for<br />

any lawful purpose. The work is made available under the Creative Commons CC0 public domain dedication.<br />

Funding: The study was supported by the NIDCR Intramural Research Program (THB), the Augustinus Foundation, Købmand Kristian Kjær og Hustrus<br />

Foundation, the Kjær-Foundation, Dagmar Marshalls Foundation, Snedkermester Sophus Jacobsen og Hustru Astrid Jacobsens Foundation, Grosserer Valdemar<br />

Foersom og Hustru Thyra Foersoms Foundation, Fabrikant Einar Willumsens Mindelegat, the Harboe Foundation (SG and LKV), the Lundbeck Foundation (KKH,<br />

SG, and LKV), the Swiss National Science Foundation 31003A-127147/1 (EH), the INSERM Avenir, Marie Curie Actions, the French National Research Agency, and<br />

the Ile-de-France Region (EC). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.<br />

Competing Interests: The authors have declared that no competing interests exist.<br />

* E-mail: thomas.bugge@nih.gov<br />

Introduction<br />

Studies conducted within the past two decades have uncovered<br />

a large family of membrane-anchored serine proteases that<br />

regulates vertebrate development, tissue homeostasis, and tissue<br />

repair by providing focal proteolysis essential for cytokine and<br />

growth factor maturation, extracellular matrix remodeling,<br />

signaling receptor activation, receptor shedding, regulation of<br />

ion channel activity, and more (reviewed in [1,2,3]). Individual<br />

members of this family regulate both vertebrate development and<br />

postnatal tissue homeostasis, including auditory and vestibular<br />

system development [4,5,6], differentiation of stratified epithelia<br />

[7,8], loss of epithelial tight junction function [9,10], failure to<br />

activate digestive enzymes [11], thyroid hormone availability [4],<br />

sodium and water homeostasis [12,13,14], iron homeostasis<br />

[15,16], and fertility [17,18]. Likewise, mounting evidence suggests<br />

that excessive or spatially dysregulated membrane-anchored serine<br />

protease activity contributes to several human disorders, including<br />

congenital malformations [19], epithelial dysfunction [20,21,22],<br />

and cancer [3].<br />

Matriptase is a modular type II transmembrane serine protease,<br />

encoded by the ST14 gene, that has pleiotropic functions in<br />

epithelial development and postnatal homeostasis, at least in part<br />

through its capacity to regulate epithelial tight junction formation<br />

in simple and stratified epithelia [2,3]. In the human and mouse<br />

epidermis, matriptase appears to function as part of a proteolytic<br />

cascade in which it acts upstream of the GPI-anchored serine<br />

protease prostasin (CAP1/PRSS8), most likely by directly activat-<br />

PLOS Genetics | www.plosgenetics.org 1 August 2012 | Volume 8 | Issue 8 | e1002937


Embryonic Cell Surface Serine Protease Cascade<br />

Author Summary<br />

Vertebrate embryogenesis is dependent upon a series of<br />

precisely coordinated cell proliferation, migration, and<br />

differentiation events. Recently, the execution of these<br />

events was shown to be guided in part by extracellular<br />

cues provided by focal pericellular proteolysis by a newly<br />

identified family of membrane-anchored serine proteases.<br />

We now show that two of these membrane-anchored<br />

serine proteases, prostasin and matriptase, constitute a<br />

single proteolytic signaling cascade that is active at<br />

multiple stages of development. Furthermore, we show<br />

that failure to precisely regulate the enzymatic activity of<br />

both prostasin and matriptase by two developmentally coexpressed<br />

transmembrane serine protease inhibitors,<br />

hepatocyte growth factor activator inhibitor-1 and -2,<br />

causes an array of developmental defects, including clefting<br />

of the embryonic ectoderm, lack of placental labyrinth<br />

formation, and inability to close the neural tube. Our study<br />

also provides evidence that the failure to regulate the<br />

prostasin–matriptase cascade may derail morphogenesis<br />

independent of the activation of known protease-regulated<br />

developmental signaling pathways. Because hepatocyte<br />

growth factor activator inhibitor–deficiency in humans is<br />

known to cause an assortment of common and rare<br />

developmental abnormalities, the aberrant activity of the<br />

prostasin–matriptase cascade identified in our study may<br />

contribute importantly to genetic as well as sporadic birth<br />

defects in humans.<br />

ing the prostasin zymogen [23,24,25,26]. Several additional<br />

candidate proteolytic substrates have been identified for matriptase<br />

in cell-based and biochemical assays, including growth factor<br />

precursors [27,28,29,30], protease-activated signaling receptors<br />

[31,32,33], ion channels [34,35], and other protease zymogens<br />

besides pro-prostasin [29,36,37]. However, the extent to which<br />

cleavage of these substrates is critical to matriptase-dependent<br />

epithelial development and maintenance of epithelial homeostasis<br />

needs to be established.<br />

Although matriptase is not required for term development in<br />

humans and most mouse strains ([24,38], and Szabo et al.,<br />

unpublished data), the membrane-anchored serine protease<br />

nevertheless is expressed in many burgeoning embryonic as well<br />

as extraembryonic epithelia [39,40,41,42]. Furthermore, we have<br />

previously shown that matriptase must be tightly regulated at the<br />

post-translational level, for successful execution of several developmental<br />

processes. Thus, loss of either of the two Kunitz-type<br />

transmembrane serine protease inhibitors, hepatocyte growth<br />

factor activator inhibitor (HAI)-1 or -2 or combined haploinsufficiency<br />

for both inhibitors, is associated with uniform embryonic<br />

lethality in mice [40,43]. Loss of HAI-1 or combined haploinsufficiency<br />

for HAI-1 and HAI-2 causes mid-gestation embryonic<br />

lethality due to failure to develop the placental labyrinth. Loss of<br />

HAI-2, in turn, is associated with three distinct phenotypes: a)<br />

Early embryonic lethality, b) mid-gestation lethality due to<br />

placental labyrinth failure, and c) neural tube defects resulting in<br />

exencephaly, spina bifida, and curly tail. All developmental defects<br />

in HAI-1- and HAI-2-deficient embryos, however, are rescued in<br />

whole or in part by simultaneous matriptase-deficiency, thus<br />

demonstrating that a matriptase-dependent proteolytic pathway is<br />

a critical morphogenic target for both protease inhibitors ([43,44],<br />

this study).<br />

In this study, we exploited the observation that HAI-1- and<br />

HAI-2-deficient mice display matriptase-dependent embryonic<br />

lethality with complete penetrance to perform a comprehensive<br />

genetic epistasis analysis aimed at identifying additional components<br />

of the matriptase proteolytic pathway. Specifically, we<br />

generated mice with simultaneous ablation of either the Spint1<br />

gene (encoding HAI-1) or the Spint2 gene (encoding HAI-2) along<br />

with genes encoding candidate matriptase targets that are coexpressed<br />

with the protease during development. We then<br />

screened for the rescue of embryonic lethality or restoration of<br />

HAI-1 and HAI-2-dependent morphogenic processes in these<br />

double-deficient mice. This analysis identified prostasin as critical<br />

to all matriptase-induced embryonic defects in both HAI-1- and<br />

HAI-2-deficient mice. Paradoxically, however, although matriptase<br />

autoactivates efficiently and prostasin is incapable of<br />

undergoing autoactivation, we found that prostasin acts upstream<br />

of matriptase in the developing embryo and is required for<br />

conversion of the matriptase zymogen to active matriptase.<br />

Finally, we explored the contribution of this newly identified<br />

prostasin-matriptase pathway to protease-activated receptor<br />

(PAR)-dependent signaling during neural tube formation [45]<br />

and now provide evidence that the pathway may be separate from<br />

the proteolytic machinery that mediates focal activation of PAR-2<br />

during neural tube closure.<br />

Results<br />

Developmental defects in HAI-2–deficient mice tightly<br />

correlate with matriptase expression levels<br />

HAI-2-deficient (Spint2 2/2 ) mice were originally reported to<br />

display embryonic lethality prior to embryonic day 8 (E8.0),<br />

presenting with severe clefting of the embryonic ectoderm at E7.5<br />

and a failure to progress to the headfold stage [44]. We previously<br />

reported, however, that approximately 50% of HAI-2-deficient<br />

mice complete early development but die at midgestation due to<br />

defective placental branching morphogenesis [43]. However, the<br />

genotyping strategy used in the latter study aimed at exploring the<br />

contribution of matriptase to this embryonic demise and only<br />

allowed for the discrimination of HAI-2-deficient mice on<br />

matriptase-sufficient (wildtype, Spint2 2/2 ;St14 +/+ , or haploinsufficient,<br />

Spint2 2/2 ;St14 +/2 ) backgrounds from a matriptase-deficient<br />

(Spint2 2/2 ;St14 2/2 ) background. Therefore, to test the possibility<br />

that early embryonic development of HAI-2-deficient mice is St14<br />

gene dosage-dependent, we first analyzed the offspring of<br />

interbred Spint2 +/2 ;St14 +/2 mice at various developmental stages.<br />

This analysis revealed that the various developmental phenotypes<br />

seen in HAI-2-deficient mice, indeed, were strongly dependent on<br />

St14 gene dosage (Figure 1A). Thus, HAI-2-deficient embryos<br />

carrying two wildtype matriptase alleles (St14 +/+ ), displayed early<br />

lethality, as evidenced by only five percent of Spint2 2/2 ;St14 +/+<br />

embryos developing beyond E9.0 and none past E10.5 (Figure 1A,<br />

blue diamonds). Inactivation of one matriptase allele (Spint2 2/2 ;<br />

St14 +/2 ), however, was sufficient to partially rescue this early<br />

embryonic lethality of HAI-2-deficient mice (Figure 1A, red<br />

squares). As reported previously [43], inactivation of both alleles of<br />

matriptase (Spint2 2/2 ;St14 2/2 ) completely restored embryonic<br />

survival and placental development and also reduced the<br />

occurrence of neural tube defects associated with the loss of<br />

HAI-2 (Figure 1A, green triangles and Table 1). Taken together,<br />

these findings show that loss of HAI-2 may lead to three distinct<br />

developmental phenotypes, dependent on the overall expression<br />

level of matriptase (Table 1): (i) early embryonic lethality occurring<br />

largely prior to E8.5, which can be partially rescued by matriptase<br />

haploinsufficiency (Spint2 2/2 ;St14 +/2 ) and completely by matriptase<br />

deficiency (Spint2 2/2 ;St14 2/2 ); (ii) placental defects resulting in<br />

mid-gestation lethality, which are observed in Spint2 2/2 ;St14 +/2<br />

embryos after E9.5, but are absent in Spint2 2/2 ;St14 2/2 embryos,<br />

PLOS Genetics | www.plosgenetics.org 2 August 2012 | Volume 8 | Issue 8 | e1002937


Embryonic Cell Surface Serine Protease Cascade<br />

Figure 1. Effect of St14 gene dosage and c-Met activity on embryonic development in HAI-1– and HAI-2–deficient mice. (A) Matriptase<br />

haploinsufficiency partially restores early embryonic development of HAI-2 deficient mice. Relative frequency of Spint2 2/2 ;St14 +/+ (blue diamonds<br />

and trend line), Spint2 2/2 ;St14 +/2 (red squares and trendline), and Spint2 2/2 ;St14 2/2 (green triangles and trendline) embryos in offspring from<br />

interbred Spint2 +/2 ;St14 +/2 mice at E8.5–E15.5. The expected 25% Mendelian frequency is shown with the dotted trend line. 59–250 embryos were<br />

genotyped at each stage. (B) Matriptase haploinsufficiency does not rescue development of HAI-1-deficient mice. Genotype distribution of E8.5–E11.5<br />

embryos and newborn (P1) offspring from interbred Spint1 +/2 ;St14 +/2 mice. No living St14 +/+ ;Spint1 2/2 or St14 +/2 ;Spint1 2/2 embryos are observed<br />

after E9.5. (C) Distribution of Spint1 genotypes in c-Met-expressing (Hgfr +/+ or Hgfr +/2 , blue bars) and c-Met-deficient (Hgfr 2/2 , green bars) embryos<br />

from interbred Spint1 +/2 ;Hgfr +/2 mice at E11.5–13.5. Loss of c-Met activity does not improve embryonic survival of HAI-1-deficient mice. (D and E)<br />

Distribution of Spint2 genotypes in c-Met-expressing (Hgfr +/+ or Hgfr +/2 , blue bars) and c-Met-deficient (Hgfr 2/2 , green bars) embryos from<br />

Spint2 +/2 ;Hgfr +/2 6Spint2 +/2 ;Hgfr +/2 (D) or Spint2 +/2 ;Hgfr +/2 6Spint2 +/2 ;Hgfr +/2 ;St14 +/2 (E) breeding pairs at E9.5–10.5. Only St14 +/2 embryos are<br />

shown in (E). Loss of c-Met does not improve survival of HAI-2-deficient embryos. (F) Frequency of exencephaly observed in 153 control<br />

(Spint2 + ;Hgfr + ), 53 c-Met- (Spint2 + ;Hgfr 2/2 ), 24 HAI-2- (Spint2 2/2 ,Hgfr + ), and 6 c-Met and HAI-2 double- (Spint2 2/2 ;Hgfr 2/2 ) deficient embryos at E9.5.<br />

Loss of c-Met activity fails to correct neural tube defects in HAI-2-deficient mice.<br />

doi:10.1371/journal.pgen.1002937.g001<br />

Table 1. Developmental defects observed in Spint1- and Spint-2-deficient mice as function of St14 expression.<br />

Genotype Phenotype Penetrance<br />

Spint1 2/2<br />

Lack of placental labyrinth, embryonic lethality<br />

at E10.5<br />

St14 +/+ St14 +/2 St14 2/2<br />

100% 100% (no rescue) 0% (complete rescue)<br />

Spint2 2/2 Early embryonic lethality at E9.5 or earlier 100% 45% (partial rescue) 0% (complete rescue)<br />

Incomplete differentiation of placental<br />

N/A 100% (no rescue) 0% (complete rescue)<br />

labyrinth, embryonic lethality at E10.5–E14.5<br />

Neural tube defects<br />

N/A<br />

Exencephaly 95–100% 18% (partial rescue; P,0.0001)<br />

Spina bifida 11% 13% (no rescue)<br />

Curly tail 89% 62% (partial rescue?; not<br />

significant)<br />

doi:10.1371/journal.pgen.1002937.t001<br />

PLOS Genetics | www.plosgenetics.org 3 August 2012 | Volume 8 | Issue 8 | e1002937


Embryonic Cell Surface Serine Protease Cascade<br />

and (iii) neural tube defects observed at or after E8.5 in most<br />

Spint2 2/2 ;St14 +/2 embryos, and partially rescued in Spint2 2/2 ;<br />

St14 2/2 embryos and term offspring.<br />

We next performed a similar analysis of the effect of St14 gene<br />

dosage on the developmental defects and embryonic lethality<br />

associated with HAI-1-deficiency by analyzing the offspring from<br />

interbred Spint1 +/2 /St14 +/2 mice (Figure 1B). As shown previously<br />

[40], a complete rescue of both the placental defects and<br />

embryonic lethality was observed in HAI-1-deficient mice<br />

expressing no matriptase (Spint1 2/2 ;St14 2/2 ). However, comparison<br />

of HAI-1-deficient mice carrying one (Spint1 2/2 ;St14 +/2 )or<br />

two (Spint1 2/2 ;St14 +/+ ) wildtype St14 alleles revealed identical<br />

defects in placental labyrinth formation and mid-gestation<br />

embryonic lethality occurring with complete penetrance<br />

(Figure 1B, Table 1, and data not shown).<br />

Activation of hepatocyte growth factor (HGF) does not<br />

contribute to placental defects in HAI-1–deficient<br />

embryos or early embryonic lethality and neural tube<br />

defects in HAI-2–deficient mice<br />

Matriptase is an efficient activator of proHGF [29,30] and<br />

dysregulated matriptase activity recently was shown to promote<br />

squamous cell carcinoma through activation of HGF-dependent c-<br />

Met signaling [46]. Furthermore, both proHGF and its cognate<br />

receptor c-Met are expressed during embryogenesis in both the<br />

placenta and the embryo [47,48]. To investigate the involvement<br />

of aberrant proHGF activation and c-Met signaling in the etiology<br />

of the defects observed in HAI-1- and HAI-2-deficient embryos,<br />

we took advantage of the fact that c-Met is only required for<br />

embryonic development beyond E13.5 [47,48]. This enabled the<br />

study of key HAI-1- and HAI-2-dependent morphogenic processes<br />

in mice homozygous for a null mutation in Hgfr (Hgfr 2/2 ),<br />

encoding c-Met. Analysis of embryos from interbred Spint1 +/2 ;<br />

Hgfr +/2 mice at E11.5–E13.5 revealed only one surviving<br />

Spint1 2/2 ;Hgfr 2/2 embryo, indicating that the loss of c-Met<br />

activity does not restore placental development or embryonic<br />

survival of HAI-1-deficient mice (Figure 1C, P,0.04, Chi-square<br />

test, and data not shown). Likewise, no Spint2 2/2 ;Hgfr 2/2<br />

embryos were detected beyond E9.5 (Figure 1D, P,0.02, Chisquare<br />

test), indicating that the inactivation of c-Met signaling does<br />

not prevent matriptase-induced early embryonic lethality in HAI-<br />

2-deficient mice. Interbreeding Spint2 +/2 ;Hgfr +/2 ;St14 +/2 mice<br />

allowed for the analysis of the impact of c-Met deficiency on the<br />

formation of neural tube defects in HAI-2-deficient mice by<br />

preventing early embryonic lethality (Figure 1E). However, all of<br />

the Spint2 2/2 ;Hgfr 2/2 ;St14 +/2 embryos isolated at E9.5 from<br />

these crosses presented with exencephaly (Figure 1F), suggesting<br />

that c-Met signaling is not critically involved in the neural tube<br />

defects caused by the absence of HAI-2. All Spint2 2/2 ;<br />

Hgfr 2/2 ;St14 +/2 embryos displayed synthetic lethality after E9.5,<br />

which precluded the direct analysis of the impact of c-Met loss on<br />

the defects in placental differentiation caused by HAI-2 deficiency<br />

(data not shown). Taken together, these findings suggest that<br />

aberrant HGF-c-Met signaling does not contribute to the<br />

matriptase-dependent defects in placentation in HAI-1-deficient<br />

embryos, or early lethality and neural tube closure of HAI-2-<br />

deficient embryos.<br />

Reduced prostasin enzymatic activity prevents<br />

developmental defects in both HAI-1– and HAI-2–<br />

deficient mice<br />

The GPI-anchored membrane serine protease, prostasin<br />

(CAP1/PRSS8), is a well-validated downstream proteolytic target<br />

for matriptase in the epidermis of mice and humans (see<br />

Introduction). To explore the possibility that matriptase acts<br />

through prostasin to cause the signature defects in embryonic<br />

development of HAI-1- and HAI-2-deficient mice, we first<br />

performed a detailed immunohistochemical analysis of prostasin<br />

expression in the developing embryo by staining histological<br />

sections from wildtype (Prss8 +/+ ) and littermate prostasin-deficient<br />

(Prss8 2/2 ) embryos with prostasin antibodies. Interestingly,<br />

prostasin was expressed in both the surface ectoderm, specifically<br />

covering the converging neuroepithelium at the time of the neural<br />

tube closure (Figure 2A and 2B, compare with 2C), and in the<br />

developing placenta, where expression was detected as early as on<br />

E8.5 and was present in the placental labyrinth in the entire period<br />

of placental differentiation (Figure 2D, 2E, 2G, and 2H, compare<br />

with 2F and 2I), thereby displaying co-expression with matriptase,<br />

HAI-1 and HAI-2 [40,41,42,43,45]. We, therefore, next directly<br />

determined the contribution of prostasin to the matriptasedependent<br />

developmental defects of HAI-1- and HAI-2-deficient<br />

mice. For this purpose, we exploited the fact that the spontaneous<br />

mutant mouse strain, frizzy, recently was described to be<br />

homozygous for a point mutation in the coding region of the<br />

Prss8 gene (Prss8 fr/fr ). This mutation results in a non-conservative<br />

V170D amino acid substitution in the prostasin protein [49].<br />

Moreover, this mutant mouse strain completes development, but<br />

displays an epidermal phenotype resembling mice carrying a<br />

hypomorphic mutation in St14 [23], suggesting reduced expression<br />

or enzymatic activity of V170D prostasin. Western blot and<br />

immunohistochemical analysis of tissues from Prss8 fr/fr mice did<br />

not reveal an obvious reduction in the level of V170D prostasin<br />

expression when compared to wildtype prostasin in Prss8 +/+<br />

littermates (data not shown). Therefore, to assess the enzymatic<br />

activity of the mutant prostasin, we generated enteropeptidaseactivated<br />

recombinant V170D prostasin, as well as enteropeptidase-activated<br />

wildtype and catalytically inactive (S238A) prostasin<br />

variants in HEK293T cells, as described previously [26]. These<br />

recombinant proteins were released from the plasma membrane<br />

by phosphatidylinositol-specific phospholipase C, activated with<br />

enteropeptidase, and their enzymatic activity towards a prostasinselective<br />

fluorogenic peptide substrate (Figure 2J) as well as their<br />

ability to form enzymatic activity-dependent covalent complexes<br />

with the serpin, protease nexin-1 (PN-1) (Figure 2K), were tested.<br />

As expected, wildtype recombinant prostasin exhibited easily<br />

detectable hydrolytic activity towards the fluorogenic peptide<br />

(Figure 2J, red line) and formed SDS-stable complexes with PN-1<br />

(Figure 2K, compare lanes 3 and 4), while prostasin not activated<br />

by enteropeptidase and the catalytically inactive S238A mutant<br />

and exhibited no detectable hydrolytic activity (Figure 2J, black<br />

and grey lines) or PN-1 binding (Figure 2K and Figure S1A, lanes<br />

2 and 12). V170D prostasin displayed a low residual enzymatic<br />

activity that was above the baseline level, as defined by the<br />

catalytically inactive S238A variant, and corresponded to about<br />

6% of the activity of wildtype prostasin (Figure 2J, blue line), while<br />

complex formation with PN-1 could not be detected (Figure 2K<br />

and Figure S1A, compare lanes 7 and 8). Taken together, these<br />

data indicated that V170D prostasin, expressed by the Prss8 fr<br />

allele, displays greatly reduced enzymatic activity. We, therefore,<br />

next interbred Spint1 +/2 ;Prss8 fr/+ and Spint1 +/2 ;Prss8 fr/fr mice and<br />

analyzed the distribution of Spint1 alleles in the newborn offspring<br />

from these crosses. Consistent with our previous findings, loss of<br />

HAI-1 was not compatible with embryonic survival of mice<br />

carrying a wildtype prostasin allele (Spint1 2/2 ;Prss8 fr/+ ) (Figure 3A,<br />

blue bars). Interestingly, however, HAI-1-deficient mice carrying<br />

two mutant prostasin alleles (Spint1 2/2 ;Prss8 fr/fr ) developed to term<br />

(Figure 3A, green bars), although they were found at a frequency<br />

PLOS Genetics | www.plosgenetics.org 4 August 2012 | Volume 8 | Issue 8 | e1002937


Embryonic Cell Surface Serine Protease Cascade<br />

Figure 2. Prostasin expression in embryonic and extraembryonic tissues. (A–C) Immunohistochemical detection of prostasin at E8.5 in<br />

epithelial cells of surface ectoderm (examples with arrows in A and B) overlying the cranial neural tube region. Specificity of staining is shown by the<br />

absence of staining of Prss8 2/2 surface ectoderm (arrow in C). Filled arrowhead shows non-specific staining of yolk sac. No expression was observed<br />

PLOS Genetics | www.plosgenetics.org 5 August 2012 | Volume 8 | Issue 8 | e1002937


Embryonic Cell Surface Serine Protease Cascade<br />

in the neuroepithelium (A and B, open arrowheads). (D–F) Immunohistochemical detection of prostasin in the chorionic ectoderm (examples with<br />

arrows) of mouse placenta at E8.5. Specificity of staining is shown by the absence of staining of Prss8 2/2 chorionic ectoderm (F). Filled arrowheads in<br />

D and F shows non-specific staining of trophoblast giant cells. No expression was detected in the trophoblast stem cell-containing chorionic<br />

epithelium (open arrowhead in E). (G–I) Immunohistochemical detection of prostasin in the placental labyrinth (examples with arrows in G and H) of<br />

mouse placenta at E12.5. Specificity of staining is shown by the absence of staining of the Prss8 2/2 labyrinth (I). No expression was detected in the<br />

trophoblast stem cell-containing chorionic epithelium (open arrowhead in H). Scale bars: A, C, D, F, G, and I, 100 mm; B, E, and H, 25 mm. (J) Enzymatic<br />

activity of wildtype (red), V170D (blue), S238A (grey), and zymogen (black) forms of prostasin. Prostasin variants were incubated with 50 mM pERTKR-<br />

AMC fluorogenic peptide at 37uC. V170D prostasin exhibited about 6% of the amidolytic activity of wildtype prostasin. No activity of catalytically<br />

inactive prostasin or prostasin zymogen was detected. (K) Western blot detection of SDS-stable complexes between prostasin and protein nexin-1<br />

(PN-1). Wildtype zymogen (lanes 1 and 2), activated wildtype (lanes 3 and 4), V170D (frizzy) zymogen (lanes 5 and 6), activated V170D (lanes 7 and 8),<br />

S238A zymogen (lanes 9 and 10), and activated S238A (lanes 11 and 12) prostasin variants were incubated with (lanes 2, 4, 6, 8, 10, and 12) or without<br />

(lanes 1, 3, 5, 7, 9, and 11) 250 ng of recombinant human PN-1. Wildtype, but not V170D or S238A variants of prostasin formed SDS-stable complexes<br />

with PN-1. Positions of pro-prostasin, activated prostasin (migrating slightly faster than the zymogen due to removal of the 12 aa propeptide that is<br />

not detected after 4–12% SDS/PAGE with anti-prostasin antibody), and prostasin/PN-1 complexes are indicated. Positions of molecular weight<br />

markers (kDa) are shown on left.<br />

doi:10.1371/journal.pgen.1002937.g002<br />

that was slightly lower than the expected Mendelian distribution<br />

(20/127, 16% vs. expected 31.75/127, 25%, P,0.05, Chi-square<br />

test). Furthermore, morphometric analysis showed that reduced<br />

prostasin activity fully restored placental labyrinth formation in<br />

HAI-1-deficient embryos, as evidenced by normal histological<br />

appearance of the labyrinth (Figure 3B–3G), thickness of the<br />

labyrinth layer (Figure 3H) and labyrinth vessel density (Figure 3I)<br />

of Spint1 2/2 ;Prss8 fr/fr embryos. Furthermore, macroscopic and<br />

histological analysis of embryos extracted between E11.5 and<br />

E13.5 failed to reveal any obvious developmental abnormalities<br />

within either embryonic or extraembryonic tissues of<br />

Spint1 2/2 ;Prss8 fr/fr mice (data not shown), and these mice were<br />

outwardly indistinguishable from their Prss8 fr/fr littermates at<br />

weaning and when followed for up to one year (Figure 3J). Taken<br />

together, these data show that the matriptase-mediated developmental<br />

defects in HAI-1-deficient mice are prostasin-dependent.<br />

To determine the impact of diminished prostasin activity on the<br />

developmental defects associated with HAI-2-deficiency, we next<br />

analyzed neural tube closure, placental differentiation, and overall<br />

survival of the offspring of interbred Spint2 +/2 ;Prss8 fr/+ mice.<br />

Analysis of the genotype distribution of embryos at E9.5–11.5 did<br />

not identify any HAI-2-deficient embryos carrying at least one<br />

wildtype Prss8 allele (Spint2 2/2 ;Prss8 +/+ or Spint2 2/2 ;Prss8 fr/+ )<br />

(Figure 4A). Interestingly, however, HAI-2-deficient embryos<br />

carrying two mutant Prss8 alleles (Spint2 2/2 ;Prss8 2fr/fr ) were found<br />

in the expected Mendelian ratio as late as E13.5–15.5 (Figure 4B,<br />

green bars). Furthermore, genotyping of newborn offspring<br />

revealed the presence of living Spint2 2/2 ;Prss8 fr/fr pups<br />

(Figure 4C, green bars), although they were found at slightly<br />

lower than expected frequency (15% vs. expected 25%, P,0.06,<br />

Chi-square test). These data strongly suggest that matriptase and<br />

prostasin act as part of a single proteolytic cascade to cause<br />

developmental defects in HAI-2-deficient mice. If this were the<br />

case, we hypothesized that lowering the activity of this cascade<br />

even further by eliminating one St14 allele from Spint2 2/2 ;Prss8 fr/fr<br />

embryos should additionally improve the term survival of HAI-2-<br />

deficient mice. Indeed, genotyping of born offspring from<br />

interbred Spint2 +/2 ;Prss8 fr/+ ;St14 +/2 mice showed a normal<br />

distribution of Spint2 alleles in Prss8 fr/fr ;St14 +/2 pups (Figure 4C,<br />

red bars), further suggesting that failure to regulate a proteolytic<br />

pathway including matriptase and prostasin accounts for all of the<br />

embryonic lethality caused by loss of HAI-2.<br />

As reported previously, neural tube defects, including exencephaly,<br />

spina bifida, and curly tail were seen in 95–100% of<br />

Spint2 2/2 ;Prss8 fr/+ ;St14 +/2 mice ([43], this study). Examination of<br />

embryonic and extraembryonic tissues from Spint2 2/2 ;Prss8 fr/fr<br />

and Spint2 2/2 ;Prss8 fr/fr ;St14 +/2 embryos, however, revealed that<br />

reduced prostasin activity sufficed to almost completely rescue the<br />

defects in both neural tube closure and placental differentiation<br />

caused by HAI-2 deficiency. Thus, macroscopic (Figure 4D and<br />

4E) and histological (Figure 4F and 4G) examination of<br />

Spint2 2/2 ;Prss8 fr/fr ;St14 +/+ and Spint2 2/2 ;Prss8 fr/fr ;St14 +/2 embryos<br />

showed that, respectively, 5% and 0%, of these embryos<br />

exhibited exencephaly when analyzed after E9.5 (Figure 4H), and<br />

no embryos with either spina bifida or curly tail were observed<br />

(data not shown). Similarly, histological analysis of placental tissues<br />

from E10.5–E13.5 Spint2 2/2 ;Prss8 fr/fr or Spint2 2/2 ;Prss8 fr/fr ;St14 +/<br />

2 embryos did not reveal any of the stereotypic defects associated<br />

with HAI-2 deficiency (Figure 4I and 4J). Thus, the overall<br />

appearance of the placental layers (Figure 4I and 4J), the thickness<br />

of the placental labyrinth (Figure 4K), and the number of fetal<br />

vessels within the labyrinth (Figure 4L), all were comparable to the<br />

HAI-2-sufficient littermate controls. In conclusion, these data<br />

document an essential role of prostasin in the etiology of all of the<br />

developmental defects previously observed in HAI-2-deficient<br />

mice.<br />

Prostasin is required for the activation of matriptase<br />

during development<br />

Matriptase was previously identified as an essential proteolytic<br />

activator of prostasin in the epidermis, and the near ubiquitous colocalization<br />

of the two membrane serine proteases in the epithelial<br />

compartment of most other adult tissues indicate that this<br />

matriptase-prostasin proteolytic pathway may be operating in<br />

multiple epithelia to maintain tissue homeostasis [23,24,25,26].<br />

The genetic epistasis analysis performed above provided strong<br />

evidence that matriptase and prostasin also are part of a single<br />

proteolytic cascade in the context of embryonic development.<br />

Furthermore, the striking overlap in expression of the two<br />

proteases documented earlier in the surface ectoderm during<br />

neural tube closure (see above) was also observed in the developing<br />

placenta (compare Figure 5A and 5B). To further investigate the<br />

functional interrelationship between the two proteases, we<br />

analyzed the levels of the activated forms of matriptase and<br />

prostasin in embryonic and placental tissues from matriptase-<br />

(St14 2/2 ) or prostasin- (Prss8 2/2 ) deficient mice at E11.5.<br />

Trypsin-like serine proteases are activated by autocatalytic or<br />

heterocatalytic cleavage after an arginine or lysine residue, located<br />

in a conserved activation motif within the catalytic domain.<br />

Activation cleavage severs the bond between the catalytic domain<br />

and upstream accessory domains, but the activated protease<br />

domain remains connected to upstream accessory domains by a<br />

disulfide bond [50]. Zymogen activation, therefore, can be<br />

detected by a mobility shift in reducing SDS-PAGE gels, which<br />

breaks the disulfide bond that keeps the two domains together.<br />

Direct detection of active matriptase in placental tissues by western<br />

blot, however, proved unsuccessful due to low signal intensity and<br />

PLOS Genetics | www.plosgenetics.org 6 August 2012 | Volume 8 | Issue 8 | e1002937


Embryonic Cell Surface Serine Protease Cascade<br />

Figure 3. Reduced prostasin activity restores placental development and embryonic survival of HAI-1–deficient mice. (A) Distribution<br />

of genotypes of born offspring of intercrossed Spint1 +/2 ;Prss8 fr/2 mice. No Spint1 2/2 mice expressing one or two wildtype prostasin alleles (Prss8 +/+ or<br />

Prss8 +/fr , blue bars) were identified, while Spint1 2/2 embryos carrying two mutant prostasin alleles (Prss8 fr/fr , green bars) were found in near-expected<br />

frequency. (B–G) Representative low (B–D) and high (E–G) magnification images showing the histological appearance of H&E-stained placental tissues<br />

of (Spint1 + ;Prss8 + ) (B and E), (Spint1 2/2 ;Prss8 + ) (C and F), and (Spint1 2/2 ;Prss8 fr/fr ) (D and G) embryos at E11.5. The thickness of the placental labyrinth<br />

(two-sided arrows between the dotted lines in B–D), as well as the number of fetal vessels (E–G, arrows) and lacunae filled with maternal blood (E–G,<br />

arrowheads) within the labyrinth is markedly reduced in prostasin-sufficient (C and F), but not in prostasin-deficient (D and G) Spint1 2/2 embryos,<br />

when compared to the controls (B and E). (H, I) Quantification of the maximum thickness of the labyrinth layer (H) and the number of fetal vessels in<br />

the placental labyrinth (I) of Spint1 + ;Prss8 + , Spint1 + ;Prss8 fr/fr , Spint1 2/2 ;Psrr8 + , and Spint1 2/2 ;Psrr8 fr/fr embryos at E11.5. The thickness of the labyrinth<br />

and fetal vessel density were strongly diminished in HAI-1-deficient mice but completely restored in HAI-1-deficient mice with low prostasin activity.<br />

(J) Outward appearance of one-year-old Spint1 2/2 ;Prss8 fr/fr and littermate Spint1 + ;Prss8 + mice. ***, p,0.0001, Student’s t-Test, two tailed. Scale bars:<br />

B–D, 100 mm; E–G, 25 mm.<br />

doi:10.1371/journal.pgen.1002937.g003<br />

a strong cross reactivity of available anti-matriptase antibodies<br />

with unrelated antigens. Similarly, direct detection of active<br />

prostasin by western blot failed due to the small difference in the<br />

electrophoretic mobility of the zymogen and the active form of the<br />

enzyme (data not shown). In order to circumvent these problems,<br />

we instead determined the amount of active matriptase and<br />

prostasin that formed inhibitor complexes with endogenous HAI-1<br />

in embryonic tissues from wildtype, matriptase-, and prostasindeficient<br />

embryos. Immunoprecipitation of protein extracts using<br />

anti-mouse HAI-1 antibodies followed by western blot with<br />

prostasin antibodies detected the presence of the 38 kDa band<br />

in the placentas of wildtype mice (Figure 5C, lanes 2 and 4) and<br />

matriptase-deficient mice (Figure 5C, lane 3). This band was not<br />

detected in placental extracts from prostasin-deficient embryos<br />

(Figure 5C, lane 1) or when anti-HAI-1 antibodies were omitted<br />

from the assay (Figure 5D, compare lanes 1 and 2), indicating that<br />

it represents the active form of prostasin released from an<br />

inhibitory complex with HAI-1. In support of this, when prostasin<br />

from either matriptase-deficient or littermate wildtype control<br />

placental tissues was released from the immunoprecipitated HAI-<br />

1-prostasin complexes by brief exposure to low pH, it was able to<br />

form SDS-stable complex with PN-1, which requires the catalytic<br />

activity of prostasin (Figure 5C, compare lane 3 with 5 and lane 4<br />

with 6). Quantification of the amount of active prostasin in<br />

PLOS Genetics | www.plosgenetics.org 7 August 2012 | Volume 8 | Issue 8 | e1002937


Embryonic Cell Surface Serine Protease Cascade<br />

Figure 4. Reduced prostasin activity restores placental differentiation, embryonic survival, and neural tube closure in HAI-2–<br />

deficient mice. (A) Distribution of Spint2 genotypes in prostasin-sufficient (Prss8 +/+ or Prss8 +/fr ) E9.5–11.5 offspring from interbred Spint2 +/2 ;Prss8 fr/+<br />

mice. No Spint2 2/2 embryos were observed (P,0.025, Chi-square test). (B) Distribution of Spint2 genotypes in prostasin-sufficient (Prss8 +/+ or Prss8 +/fr ,<br />

blue bars) and prostasin-deficient (Prss8 fr/fr , green bars) mouse embryos from interbred Spint2 +/2 ;Prss8 fr/+ mice at E13.5–15.5. No prostasin-expressing<br />

Spint2 2/2 embryos were observed (P,0.001, Chi-square test), while survival of prostasin-deficient Spint2 2/2 embryos was restored. (C) Distribution of<br />

Spint2 genotypes in newborn prostasin-sufficient, matriptase wildtype (Prss8 +/+ or Prss8 +/fr ;St14 +/+ , blue bars), prostasin-deficient, matriptase wildtype<br />

(Prss8 fr/fr ;St14 +/+ , green bars), and prostasin-deficient, matriptase haploinsufficient (Prss8 fr/fr ;St14 +/2 , red bars) offspring from Spint2 +/2 ;Prss8 fr/2<br />

6Spint2 +/2 ;Prss8 fr/+ ;St14 +/2 breeding pairs. Reduced prostasin activity restored embryonic survival of Spint2 2/2 mice partially in matriptase wildtype<br />

and completely in matriptase haploinsufficient mice. (D–G) Macroscopic (D and E) and histological (H&E staining) (F and G) appearance of the HAI-2-<br />

deficient, matriptase- and prostasin-sufficient (Spint2 2/2 ;Prss8 +/+ or Prss8 +/fr , St14 +/2 , D and F) or HAI-2- and prostasin-deficient, matriptase-sufficient<br />

(Spint2 2/2 ;Prss8 fr/fr , St14 +/+ or St14 +/2 ) (E and G) embryos at E9.5. HAI-2 deficiency prevents convergence of neural folds in the cranial region of neural<br />

tube (D and F, arrows) leading to exencephaly. Convergence and fusion of neural folds are restored in HAI-2-deficient mice with low prostasin activity<br />

(E and G, arrows). Presence of medial (F, open arrowhead) and absence of dorsolateral (F, arrowheads) hinge points. (H) Frequency of exencephaly in<br />

E9.5–18.5 Spint2 2/2 embryos with different levels of prostasin activity (Prss8 +/+ , Prss8 fr/+ or Psrr8 fr/fr ) and matriptase (St14 +/+ , St14 +/2 or St14 2/2 ). The<br />

frequency of neural tube defects is inversely correlated with the combined number of wildtype Prss8 and St14 alleles. A total of 524 embryos were<br />

analyzed. (I–L) Histological appearance (H&E staining) (I and J), thickness of placental labyrinth (K), and number of fetal vessels within the labyrinth (L)<br />

in the placentas of HAI-2 and prostasin-sufficient (Spint2 + ;Prss8 + ) and HAI-2 and prostasin double-deficient (Spint2 2/2 ;Prss8 fr/fr ) embryos at E12.5.<br />

Reduced prostasin activity restores differentiation of placental labyrinth in Spint2 2/2 mice to levels not significantly (N.S.) different from wildtype<br />

littermate controls. Arrows in I and J show examples of fetal vessels. Scale bars: F, 50 mm G, I, and J, 100 mm.<br />

doi:10.1371/journal.pgen.1002937.g004<br />

PLOS Genetics | www.plosgenetics.org 8 August 2012 | Volume 8 | Issue 8 | e1002937


Embryonic Cell Surface Serine Protease Cascade<br />

Figure 5. Prostasin is required for the activation of matriptase during placental differentiation. (A and B) Expression of prostasin (A) and<br />

matriptase (B) in placental tissues of wildtype mice at E11.5. Both proteins were expressed in the chorionic (arrows) and labyrinthine (arrowheads)<br />

trophoblasts. (C) Western blot detection of active prostasin in the fetal part of the placenta of wildtype (Prss8 +/+ and St14 +/+ , lanes 2, 4, and 6),<br />

prostasin-deficient (St14 +/+ ;Prss8 2/2 ) (lane 1), and matriptase-deficient (St14 2/2 ;Prss8 +/+ ) (lanes 3 and 5) embryos at E11.5 after immunoprecipitation<br />

with anti-mouse HAI-1 antibodies. Immunoprecipitated proteins in lanes 5 and 6 were acid-exposed to dissociate prostasin-HAI-1 complexes, and<br />

then incubated with PN-1 prior to western blot analysis. Positions of bands corresponding to active prostasin, prostasin/PN-1 complex, as well as nonspecific<br />

signals of IgG heavy and light chains are indicated on the right. Positions of molecular weight markers (kDa) are shown on the left. (D)<br />

Omission of anti-HAI-1 antibody resulted in loss of detectable prostasin (compare lanes 1 and 2), indicating that the detected prostasin formed<br />

complexes with HAI-1. (E) Quantification of the relative amount of active prostasin in wildtype and matriptase placentae by densitometric scanning of<br />

prostasin western blots of HAI-1 immunoprecipitated material from (Prss8 2/2 ;St14 +/+ ,N=3,Prss8 +/+ ;St14 2/2 , N = 3, and Prss8 +/+ ;St14 +/+ , N = 6). Data<br />

are shown as mean 6 standard deviation (N.S., not significant). (F and G) Western blot detection of active matriptase in the fetal part of the placenta<br />

at E11.5 (F) after anti-HAI-1 immunoprecipitation, and in the epidermis of newborn skin (G) of wildtype (Prss8 +/+ and St14 +/+ ) (F, lanes 1, and 3, and G,<br />

lane 2), prostasin-deficient (St14 +/+ ;Prss8 2/2 ) (F, lane 4 and G, lane 1), and matriptase-deficient (St14 2/2 ;Prss8 +/+ ) (F, lane 2, and G, lane 3) embryos. A<br />

30 kDa band representing the active serine protease domain of matriptase (Mat SPD) was present in extracts from wildtype (lanes 1 and 3 in F), but<br />

not in matriptase- (lane 2 in F) or prostasin-deficient (lane 4 in F) placenta. Zymogen (Mat FL) and active (Mat SPD) forms of matriptase were detected<br />

in extracts from both wildtype and prostasin-deficient, but not matriptase-deficient epidermis. (H–H0) Immunohistochemical staining of matriptase in<br />

control Prss8 + (H) and prostasin-deficient Prss8 2/2 (H9) placenta at E11.5. Specificity of staining of chorionic and labyrinthine trophoblasts (examples<br />

with arrows) is shown by the absence of staining of corresponding cells in St14 2/2 placenta (H0). Insets in H and H9 are parallel sections stained with<br />

prostasin antibodies. Open arrowheads in H–H0 show examples of non-specific staining. Scale bars: A, B, H, H9, and H0, 50mm.<br />

doi:10.1371/journal.pgen.1002937.g005<br />

PLOS Genetics | www.plosgenetics.org 9 August 2012 | Volume 8 | Issue 8 | e1002937


Embryonic Cell Surface Serine Protease Cascade<br />

wildtype and prostasin-deficient placentae by densitometric scans<br />

of western blots showed that the loss of matriptase did not affect<br />

the amount of active prostasin (Figure 5E). Taken together, these<br />

data suggest that the developing placenta does not require<br />

matriptase for the activation of prostasin.<br />

Detection of matriptase by western blot after immunoprecipitation<br />

with anti-HAI-1 antibodies revealed the presence of a<br />

30 kDa band corresponding to the activated matriptase serine<br />

protease domain in wildtype placental tissues, but not matriptasedeficient<br />

placental tissues (Figure 5F, compare lanes 1 and 2).<br />

Surprisingly, however, the active form of matriptase was also<br />

absent in the extracts from prostasin-deficient placentae (Figure 5F,<br />

compare lanes 3 and 4). The absence of matriptase was observed<br />

in four independent experiments using placentae from a total of<br />

seven prostasin-deficient mice and their prostasin-sufficient littermate<br />

controls (Figure S1B and data not shown). As expected,<br />

analysis of skin extracts from prostasin-deficient newborn mice and<br />

wildtype littermate controls using the same western blot conditions<br />

clearly showed the presence of the active form of matriptase in<br />

both the control and prostasin-deficient mice (Figure 5G, compare<br />

lanes 1 and 2 with lane 3), demonstrating that differences in the<br />

functional relationship between the two proteases exist in different<br />

tissues. Immunohistochemistry of placentae from littermate<br />

control and prostasin-deficient embryos showed no obvious<br />

difference in levels or pattern of matriptase expression (compare<br />

Figure 5H and 5H9).<br />

To further substantiate the above findings, we next determined<br />

if prostasin could serve as an activator of matriptase in a<br />

reconstituted cell-based assay. For this purpose, we transiently<br />

transfected HEK-293 cells with expression vectors encoding HAI-<br />

1 (to allow for efficient matriptase expression) and wildtype or<br />

catalytically inactive matriptase. The transfected cells were then<br />

exposed to soluble recombinant prostasin or vehicle, and<br />

matriptase activation was analyzed six hours later by western blot<br />

of cell lysates (Figure 6A) or conditioned medium (Figure 6B).<br />

Interestingly, soluble prostasin efficiently activated matriptase, as<br />

evidenced by the large increase in the amount of the liberated<br />

matriptase serine protease domain (Mat SPD, Figure 6A and 6B,<br />

compare lanes 1 and 2) after reducing SDS/PAGE, and a<br />

corresponding diminution of the amount of matriptase zymogen<br />

(Mat SEA, Figure 6A and 6B, compare lanes 1 and 2). Activation<br />

site cleavage of matriptase by prostasin did not require matriptase<br />

catalytic activity, as shown by the increased amount of the isolated<br />

matriptase serine protease domain in prostasin-treated cells<br />

expressing a catalytically inactive matriptase (Figure 6A and 6B,<br />

compare lanes 3 and 4). Similar results were obtained when<br />

matriptase-transfected HEK-293 cells were transfected with a<br />

prostasin expression vector, rather than being treated with soluble<br />

prostasin (data not shown). To investigate if the prostasin-activated<br />

matriptase displayed functional activity, the HEK-293 cells<br />

described above were also transfected with a PAR-2 expression<br />

vector and a serum response element (SRE)-luciferase reporter<br />

plasmid to measure PAR-2 activity (Figure 6C). Exposure of<br />

serum-starved cells to soluble prostasin resulted in a large increase<br />

in luciferase activity in cells transfected with wildtype matriptase<br />

(Figure 6C, left panels), but not in cells transfected with<br />

catalytically inactive matriptase (Figure 6C, second panels from<br />

left), with HAI-1 alone (Figure 6C, second panels from right) or<br />

with empty vector (Figure 6C, right panels). Taken together, the<br />

data indicate that prostasin can proteolytically activate matriptase<br />

and is critical for the generation of active matriptase during<br />

placental development. Detection of active matriptase and<br />

prostasin in the embryo by western blot or by anti-HAI-1<br />

immunoprecipitation failed to detect either of the proteases, likely<br />

due to the restricted expression of both proteins (data not shown).<br />

Developmental defects in HAI-2–deficient embryos are<br />

not caused by aberrant activity of the epithelial sodium<br />

channel<br />

Both matriptase and prostasin have been reported to activate<br />

the epithelial sodium channel (ENaC) in cell-based assays, and<br />

prostasin is a critical regulator of ENaC activity during alveolar<br />

fluid clearance in mouse lungs and likely regulates ENaC activity<br />

in many other adult organs [51,52]. Immunohistological analysis<br />

of embryonic tissues at E11.5 revealed strong expression of ENaC<br />

in the developing labyrinth layer of the placenta (Figure 6D). No<br />

ENaC expression was detected in the embryo proper (data not<br />

shown). To investigate a possible involvement of ENaC in the<br />

etiology of prostasin-matriptase-induced developmental defects in<br />

HAI-2-deficient mice, pregnant females from Spint2 +/2 mice bred<br />

to Spint2 +/2 ;St14 +/2 mice were treated daily between E5.5–8.5<br />

with the pharmacological inhibitor of ENaC activity, amiloride,<br />

which is known to cross the feto-maternal barrier [53]. Genotyping<br />

of embryos extracted at E9.5 from these crosses did not<br />

identify any Spint2 2/2 ;St14 +/+ embryos (Figure S1C), indicating<br />

that the inhibition of ENaC activity is not sufficient to prevent<br />

early embryonic lethality resulting from the loss of HAI-2.<br />

Furthermore, all of the seven Spint2 2/2 ;St14 +/2 embryos identified<br />

in this experiment exhibited exencephaly, suggesting that<br />

ENaC activity is not critically involved in the etiology of neural<br />

tube defects in HAI-2-deficient mice (Figure 6E). Similarly, genetic<br />

inactivation of the a subunit of ENaC (encoded by Scnn1a), which<br />

is necessary for channel activity in vivo [54,55], failed to rescue<br />

embryonic development of HAI-2-deficient animals, as evidenced<br />

by a complete absence of any surviving Spint2 2/2 ;Scnn1a 2/2<br />

double-deficient embryos at E9.5 from Spint2 +/2 ;Scnn1a +/2 mice<br />

bred to Spint2 +/2 ;Scnn1a +/2 mice (Figure 6F) and the failure of<br />

Spint2 2/2 ;Scnn1a 2/2 double-deficient mice to appear in the<br />

newborn offspring from Spint2 +/2 ;Scnn1a +/2 mice bred to<br />

Spint2 +/2 ;Scnn1a +/2 ;St14 +/2 mice (Figure 6G). Taken together,<br />

these data do not support the critical involvement of aberrant<br />

ENaC activity in the developmental defects resulting from lack of<br />

HAI-2 regulation of the prostasin-matriptase proteolytic pathway.<br />

Excess PAR-2 signaling does not cause developmental<br />

defects in HAI-2–deficient mice<br />

Matriptase and prostasin are co-expressed with PAR-2 in<br />

surface ectoderm during neural tube closure ([43,45], this study),<br />

and matriptase displays extraordinarily favorable activation<br />

kinetics towards PAR-2 in cell-based assays [43,45]. Furthermore,<br />

activation of PAR-2 (encoded by the F2rl1 gene) was recently<br />

shown to contribute to neural tube closure (see below). These data<br />

suggested that some, or all, of the prostasin- and matriptasedependent<br />

defects in HAI-2-deficient mice could be caused by<br />

excess PAR-2 signaling. To test this hypothesis, we interbred<br />

Spint2 +/2 ;F2rl1 +/2 mice and genotyped the ensuing embryos at<br />

E9.5. This analysis failed to identify any Spint2 2/2 ;F2rl1 2/2<br />

embryos (Figure 7A). Thus, the loss of PAR-2 activity is not<br />

sufficient to overcome matriptase- and prostasin-dependent early<br />

embryonic lethality in HAI-2-deficient mice. When the early<br />

embryonic survival was improved by matriptase haploinsufficiency<br />

(see above), analysis of neural tubes at E9.5 revealed exencephaly<br />

in 100% of Spint2 2/2 ;F2rl1 2/2 St14 +/2 embryos, identical to the<br />

frequency of defects observed in littermate HAI-2-deficient<br />

embryos expressing PAR-2 (Spint2 2/2 ;F2rl1 +/+ or F2rl1 +/2 ;<br />

St14 +/2 ) (Figure 7B). Thus, excess PAR-2 activation does not<br />

PLOS Genetics | www.plosgenetics.org 10 August 2012 | Volume 8 | Issue 8 | e1002937


Embryonic Cell Surface Serine Protease Cascade<br />

Figure 6. Prostasin activates matriptase on the surface of HEK293 cells. (A and B) Western blot detection of matriptase in cell lysates (A) and<br />

in the conditioned medium (B) from HEK293 cells transiently transfected with wildtype recombinant human matriptase and HAI-1 expression vectors<br />

(lanes 1 and 2), catalytically inactive (S805A) matriptase and HAI-1 (lanes 3 and 4), HAI-1 alone (lanes 5 and 6), and cells transfected with a control<br />

empty vector (lanes 7 and 8) that were incubated with (lanes 2, 4, 6, and 8) or without (lanes 1, 3, 5, and 7) 100 nM soluble recombinant human<br />

prostasin. Addition of prostasin promoted conversion of the matriptase zymogen to its activated two-chain form. Positions of bands corresponding<br />

to full length matriptase (Mat FL), matriptase pro-enzyme processed by autocatalytic cleavage within the SEA domain (Mat SEA), and activated<br />

matriptase serine protease domain (Mat SPD) are indicated on the right. Positions of molecular weight markers (kDa) are shown on left. (C)<br />

PLOS Genetics | www.plosgenetics.org 11 August 2012 | Volume 8 | Issue 8 | e1002937


Embryonic Cell Surface Serine Protease Cascade<br />

Quantification of the activation of PAR-2 in HEK293 cells expressing recombinant human PAR-2 in combination with wildtype (WT) or inactive (S805A)<br />

variants of matriptase and HAI-1, HAI-1 alone, or transfected with an empty vector, incubated without (blue bars) or with (red bars) 100 nM soluble<br />

recombinant human prostasin. Prostasin induced matriptase activity-dependent activation of PAR-2. (D) Immunohistochemical analysis of the<br />

expression of the gamma subunit of the epithelial sodium channel (ENaC) in placenta of control mice at E11.5. The expression was detected in the<br />

populations of chorionic (arrow) and labyrinthine (arrowhead) trophoblasts. Scale bar: 50 mm. (E) Frequency of exencephaly in amiloride-treated<br />

wildtype (Spint2 +/+ , N = 56), untreated HAI-2-deficient (Spint2 2/2 , N = 12) and amiloride-treated HAI-2-deficient (Spint2 2/2 , N = 7) embryos at E9.5.<br />

Amiloride treatment failed to rescue neural tube defects in Spint2 2/2 ; St14 +/2 embryos. (F) Distribution of Spint2 genotypes in ENaC-expressing<br />

(Scnn1a +/+ or Scnn1a +/2 , blue bars) and ENaC-deficient (Scnn1a 2/2 , green bars) offspring from Spint2 +/2 ;Scnn1a +/2 6Spint2 +/2 ;Scnn1a +/2 breeding<br />

pairs at E9.5. Loss of ENaC expression did not rescue early embryonic lethality in Spint2 2/2 mice. (G) Distribution of Spint2 genotypes in matriptasehaploinsufficient<br />

ENaC-expressing (St14 +/2 ;Scnn1a +/+ or Scnn1a +/2 , blue bars) and ENaC-deficient (St14 +/2 ; Scnn1a 2/2 , green bars) offspring from<br />

Spint2 +/2 ;Scnn1a +/2 , St14 +/2 6Spint2 +/2 ;Scnn1a +/2 ;St14 +/+ breeding pairs at birth. Loss of ENaC expression did not rescue overall embryonic survival in<br />

Spint2 2/2 ; St14 +/2 mice.<br />

doi:10.1371/journal.pgen.1002937.g006<br />

appear to be critically involved in the etiology of neural tube<br />

defects in HAI-2-deficient mice.<br />

Loss of matriptase does not cause neural tube defects in<br />

PAR-1–deficient embryos<br />

Rac1 activation in surface ectoderm through G i , initiated by<br />

either PAR-1 or PAR-2 activation was recently shown to be<br />

required for neural tube closure. Thus, mice with combined, but<br />

not single, deficiency in PAR-1 and PAR-2 display exencephaly<br />

with high frequency [45]. As matriptase and prostasin are coexpressed<br />

with PAR-2 in surface ectoderm during neural tube<br />

closure ([43,45], this study), we next investigated if the prostasinmatriptase<br />

cascade identified in the current study contributes to<br />

physiological PAR-2 activation during neural tube closure. To test<br />

this, we generated mice with combined deficiency in PAR-1<br />

(encoded by the F2r gene) and matriptase (F2r 2/2 ;St14 2/2 ). If<br />

matriptase was essential for the activation of PAR-2 during neural<br />

tube closure, these mice should phenocopy mice with a combined<br />

PAR-1 and PAR-2 deficiency, including embryonic lethality and<br />

high susceptibility to cranial neural tube defects [45]. However,<br />

analysis of the midgestation embryos from intercrossed<br />

F2r +/2 ;St14 +/2 mice yielded the expected distribution of all<br />

genotypes (Figure 7C). Furthermore, none of 20 observed F2r 2/<br />

2 ;St14 2/2 mice displayed cranial neural tube defects, although the<br />

PAR-1 deficiency alone or in combination with matriptase<br />

deficiency occasionally led to defects in the closure of the posterior<br />

neural tube, resulting in spina bifida and curly tail (Figure 7D).<br />

HAI-2 and PAR-1/PAR-2 regulate different stages of<br />

neural tube development<br />

The lack of functional interaction between prostasin-matriptase<br />

and PAR-1/PAR-2 regulated signaling pathways evidenced from<br />

the above experiments suggested that the two pathways are either<br />

involved in two essential, non-redundant mechanisms regulating<br />

the same steps of neural tube closure, or that they may regulate<br />

different stages of the process. To distinguish between the two<br />

possibilities, we performed a detailed morphologic comparison of<br />

the neural tube defects caused by loss of HAI-2 and by the<br />

combined loss of PAR-1 and PAR-2. Macroscopic analysis showed<br />

significant differences in the types of neural tube defects in the two<br />

mutant mouse strains. In PAR-1 and PAR-2 double-deficient<br />

mice, the defects were almost exclusively restricted to the<br />

hindbrain region of the cranial neural tube (Figure 7E and<br />

Table 2). In addition, PAR-1 and PAR-2 double-deficient embryos<br />

did not exhibit any obvious abnormalities during the early stages<br />

of neural tube closure, as all of the embryos analyzed before E10.5<br />

showed normal elevation and conversion of the opposing neural<br />

folds, as well as the completion of the neural fold fusion at initial<br />

closure points 1 and 2 at hindbrain/cervix and forebrain/<br />

midbrain boundaries, respectively (compare Figure 7F with 7G,<br />

and 7F9 with 7G9). As a result, the neural tube defects in these<br />

mice were generally only obvious after E10.5 (compare Figure 7H<br />

with 7I and 7J), and were generally restricted to the hindbrain<br />

region of the neural tube, with less than five percent exhibiting<br />

exencephaly that extended to the midbrain region (Figure 7I and<br />

Table 2). In contrast, HAI-2 deficiency was generally associated<br />

with a failure of cranial neural tube closure that was obvious at<br />

E9.5 or earlier, and was due to the inability of neural folds to<br />

elevate properly, and to come into juxtaposition necessary for the<br />

fusion (compare Figure 7F with 7K, and 7F9 with 7K9). The fusion<br />

at closure point 1 of HAI-2-deficient mice was completed in all<br />

embryos analyzed at E9.5 or later, and no case of craniorachischisis<br />

was observed (Table 2). However, 10 percent of Spint2 2/2<br />

embryos failed to initiate fusion at closure point 2, resulting in<br />

exencephaly that extended from forebrain region to the hindbraincervical<br />

boundary (Figure 7K and 7K9). In addition, even in the<br />

embryos that successfully initiated the fusion at closure point 2, the<br />

exencephaly was more extensive than the one observed in PAR-1<br />

and PAR-2 double-deficient embryos, typically spanning the entire<br />

midbrain and hindbrain regions (compare Figure 7L and 7I).<br />

Finally, histological analysis of affected embryos at E9.5 showed<br />

that dorsolateral hinge points (DLHPs) critical for the final stages<br />

of the neural tube closure were absent in 97 percent of Spint2 2/2<br />

embryos (Figure 4D and 4F, Figure 7K and 7K9, and Table 2),<br />

while F2r 2/2 ;F2rl1 2/2 embryos generally exhibited DLHP<br />

formation indistinguishable from wildtype littermate controls<br />

(Figure 7G, 7G9, and 7M, compare to Figure 4E, 4G, Figure 7F<br />

and 7F9). Thus, substantial differences are observed in the<br />

location, frequency, extent, and onset of the neural tube defects<br />

of HAI-2-deficient mice and PAR-1 and PAR-2 double-deficient<br />

mice, further indicating the independent roles of, respectively,<br />

repression and activation of the two protease-regulated pathways<br />

in distinct stages of neural tube formation.<br />

Discussion<br />

In this study, we exploited the uniform matriptase-dependent<br />

embryonic lethality of mice deficient in hepatocyte growth factor<br />

activator inhibitors as a means to genetically identify novel<br />

molecules and pathways regulating and being regulated by<br />

matriptase in the developing embryo by epistasis analysis. This<br />

analysis resulted in a number of unexpected findings. First, we<br />

found that prostasin is an essential component of the matriptasedependent<br />

molecular machinery that causes early embryonic<br />

lethality, derails placental labyrinth formation, and causes defects<br />

in neural tube closure in these mice. This shows that both proteins<br />

are expressed, are active, functionally interact, and must be<br />

regulated by hepatocyte growth factor activator inhibitors already<br />

during early development. Surprisingly, however, rather than<br />

being a downstream effector of matriptase function, as previously<br />

established for both mouse and human epidermis ([23,24,25,26],<br />

this study), prostasin acts upstream of matriptase during embryogenesis<br />

and is essential for activation of the matriptase zymogen.<br />

PLOS Genetics | www.plosgenetics.org 12 August 2012 | Volume 8 | Issue 8 | e1002937


Embryonic Cell Surface Serine Protease Cascade<br />

Figure 7. Neural tube defects and embryonic lethality in HAI-2–deficient mice are not dependent on PAR-2, and combined PAR-1<br />

and matriptase deficiency does not phenocopy combined PAR-1 and PAR-2 deficiency. (A) Distribution of Spint2 genotypes at E9.5 in<br />

PAR-2-expressing (F2rl1 +/+ or F2rl1 +/2 , blue bars) and PAR-2-deficient (F2rl1 2/2 , green bars) offspring from interbred Spint2 +/2 ,F2rl1 +/2 mice. No<br />

Spint2 2/2 embryos were detected irrespective of PAR-2 expression. (B) Frequency of exencephaly observed in HAI-2 and PAR-2-sufficient<br />

(Spint2 + ;F2rl1 + N = 366), PAR-2-deficient (Spint2 + ;F2rl1 2/2 , N = 164), HAI-2-deficient (Spint2 2/2 ,F2rl1 + , N = 18), and PAR-2 and HAI-2 double-<br />

(Spint2 2/2 ;F2rl1 2/2 , N = 12) deficient embryos extracted at E9.5–E11.5. Loss of PAR-2 activity fails to correct neural tube defects in HAI-2-deficient<br />

embryos. (C) Distribution of St14 alleles at E11.5–15.5 in PAR-1-expressing (F2r +/+ or F2r +/2 , blue bars) and PAR-1-deficient (F2r 2/2 , green bars)<br />

embryos from interbred St14 +/2 ;F2r +/2 mice. Loss of PAR-1 activity does not affect embryonic survival of matriptase-deficient mice. (D) Frequency of<br />

exencephaly (Ex), spina bifida (SB), and curly tail (CT) in E9.5–18.5 embryos with different levels of expression of PAR-1 (F2r + or F2r 2/2 ) and matriptase<br />

(St14 + or St14 2/2 ). A total of 326 embryos were analyzed. Loss of matriptase does not significantly increase the incidence of neural tube defects in<br />

PAR-1-deficient embryos. (E) Comparison of the severity of exencephaly in HAI-2-deficient (Spint2 2/2 , N = 29) and PAR-1 and PAR-2 double-deficient<br />

(F2r 2/2 ;F2rl1 2/2 , N = 39) embryos. 95% of affected F2r 2/2 ;F2rl1 2/2 embryos exhibited exencephaly that was confined to hindbrain region of the<br />

PLOS Genetics | www.plosgenetics.org 13 August 2012 | Volume 8 | Issue 8 | e1002937


Embryonic Cell Surface Serine Protease Cascade<br />

cranium (HB only, green bars), with the remaining 5% extending to the midbrain region (MB-HB, blue bars). In contrast, only 10% of exencephalies<br />

observed in Spint2 2/2 -deficient mice were confined to the hindbrain, with 59% extended to midbrain, and 31% to forebrain region (FB-HB, red bars).<br />

(F–G9) Ventral (F and G) and dorsal (F9 and G9) view of non-affected control (F and F9) and affected PAR-1 and PAR-2 double-deficient (F2r 2/2 ;F2rl1 2/<br />

2 ) (G and G9) embryos at E9.5. The initial stages of neural tube closure all appear to be unaffected by the combined absence of PAR-1 and PAR-2. (H–<br />

J) Appearance of control (H) and PAR-1 and PAR-2 double-deficient embryos with exencephaly (I and J) at E14.5. Exencephaly in 95% of the affected<br />

PAR-1 and PAR-2 double-deficient embryos was restricted to hindbrain region (HB, two-sided arrow in I) and extended to midbrain (MB-HB, two-sided<br />

arrow in J) in only 5% of the cases. (K and K9) Ventral (K) and dorsal (K9) view of the macroscopic appearance of HAI-2-deficient (Spint2 2/2 ) embryos at<br />

E9.5. Divergence of neural folds (arrows) and defects in neural tube closure extending from forebrain region to cervix are obvious. Open arrowheads<br />

show normal formation of medial hinge points. (L) Macroscopic appearance of a HAI-2-deficient embryo with exencephaly at E14.5. 90% of embryos<br />

presented with exencephaly that included at least midbrain and hindbrain regions of the developing cranium. (M) Histological appearance (nuclear<br />

fast red staining) of PAR-1 and PAR-2 double-deficient embryo with exencephaly at E9.5. Defined medial (arrow) and dorsolateral (arrowheads) hinge<br />

points are clearly visible. Scale bar: 150 mm.<br />

doi:10.1371/journal.pgen.1002937.g007<br />

This finding is perplexing, as matriptase is well-established to be<br />

able to auto-activate, as most clearly evidenced by the inability of<br />

recombinant matriptase protein with the catalytic triad serine<br />

mutated to alanine (S805A matriptase) to undergo activation site<br />

cleavage [56]. Furthermore, prostasin shows no catalytic activity<br />

towards peptide sequences derived from the prostasin pro-peptide<br />

[57] and no reports of prostasin auto-activation have appeared to<br />

date. Substantiating this finding, however, we found that prostasin<br />

efficiently activated the matriptase zymogen in a reconstituted cellbased<br />

assay. These findings are aligned with a recent study<br />

showing that PAR-2 activation in some cultured cells, caused by<br />

exposure of cultured cells to exogenously added activated<br />

prostasin, was blunted by a neutralizing antibody directed against<br />

matriptase [45], providing further evidence that complex and<br />

context-specific relationships between the two membrane-anchored<br />

serine proteases may exist in vivo. Another important<br />

finding relating to the developmental prostasin-matriptase cascade<br />

identified in this study emanated from our biochemical analysis of<br />

placental tissues, which revealed that activated forms of both<br />

matriptase and prostasin were present in a complex with HAI-1 in<br />

placental tissues. This indicates that the regulation of the prostasinmatriptase<br />

cascade by HAI-1 (and likely HAI-2) may occur by<br />

controlling both prostasin and matriptase proteolytic activity.<br />

Furthermore, as both HAI-1 and HAI-2 are very promiscuous and<br />

display potent inhibitory activity towards a number of trypsin-like<br />

serine proteases in vitro [58,59,60,61,62,63,64], it is throughout<br />

plausible that they may also regulate the activity of as yet<br />

unidentified proteases that act upstream of, downstream of or<br />

between prostasin and matriptase. Such profound complexities in<br />

zymogen activation relationships between trypsin-like serine<br />

proteases and for the promiscuity of their cognate inhibitors have<br />

Table 2. Comparison of morphologic features of neural tube<br />

defects observed in Spint2 2/2 and F2r 2/2 ; F2rl1 2/2 mice.<br />

Spint2 2/2 F2r 2/2 ; F2rl1 2/2<br />

Process<br />

Formation of medial hinge point 100% (28/28) 100% (38/38)<br />

Formation of dorsolateral hinge points 3% (1/31) 97% (28/29)<br />

Completion of C1 fusion 100% (28/28) 100% (66/66)<br />

Completion of C2 fusion 90% (27/30) 100% (66/66)<br />

Extent of neural tube defect<br />

Hindbrain only 10% (3/29) 95% (37/39)<br />

Hindbrain and midbrain 59% (17/29) 5% (2/39)<br />

Forebrain to cervix 31% (9/29) 0% (0/39)<br />

Craniorachischisis 0% (0/29) 0% (0/39)<br />

doi:10.1371/journal.pgen.1002937.t002<br />

long been recognized in the coagulation, fibrinolytic, complement,<br />

and digestive systems. The current findings, thus, serve to<br />

underscore that our knowledge of the molecular workings of<br />

membrane-anchored serine proteases is still fragmentary, due to<br />

their quite recent emergence as a protease subfamily.<br />

The outcome of our epistasis analysis querying the contribution<br />

of proHGF, PAR-2, and ENaC to the prostasin and matriptasedependent<br />

embryonic demise of HAI-1- and HAI-2-deficient mice<br />

also was unanticipated. Each of the three proteins has been<br />

genetically validated as a substrate for either matriptase or<br />

prostasin in developmental or post-developmental processes, has<br />

established functions in embryonic development, and is developmentally<br />

co-expressed with both proteases. Nevertheless, their<br />

genetic elimination failed to prevent or alleviate any of the<br />

abnormalities caused by the loss of HAI-1 or HAI-2. Importantly,<br />

our analysis does not exclude that cleavage of either of the three<br />

proteins must be suppressed by HAI-1 or HAI-2 at later stages of<br />

development that cannot be analyzed by the current experimental<br />

approach. Also, the possibility that the lethality of HAI-1- or HAI-<br />

2-deficient embryos is caused by the simultaneous cleavage of<br />

more than one of these substrates cannot be formally excluded. It<br />

was particularly surprising that the neural tube defects associated<br />

with HAI-2-deficiency were unrelated to either excessive or<br />

reduced (through desensitization) PAR-2 activity, despite the<br />

unequivocal contribution of PAR-2 signaling to neural tube<br />

closure, and the wealth of strong circumstantial evidence that<br />

prostasin and matriptase contribute to PAR-2 activation in this<br />

process [45,65]. Equally surprising in this regard, the combined<br />

loss of PAR-1 and matriptase failed to cause the neural tube<br />

closure defects observed in PAR-1 and PAR-2 double-deficient<br />

embryos, showing that matriptase is not essential for initiation of<br />

physiological PAR-2 signaling during neural tube formation.<br />

Previous analysis has identified five other membrane-anchored<br />

serine proteases and fourteen secreted trypsin-like serine proteases<br />

that are expressed during neural tube formation, some of which<br />

can activate PAR-2 in cell-based assays [43,45]. It is therefore<br />

possible that the prostasin-matriptase cascade does contribute to<br />

PAR-2 activation during neural tube closure, but sufficient<br />

residual activation of PAR-2 by other developmentally coexpressed<br />

serine proteases takes place in its absence to allow for<br />

completion of this developmental process. Nevertheless, the careful<br />

comparison of the morphology of neural tube defects in PAR-1<br />

and PAR-2, and HAI-2-double deficient embryos performed here<br />

revealed distinct differences in terms of their anatomical location<br />

and the stage of developmental failure. Taken together, these data<br />

suggest that promotion of neural tube closure by HAI-2<br />

suppression of the prostasin-matriptase cascade and promotion<br />

of neural tube closure by PAR-1/PAR-2 signaling may be<br />

temporally and spatially distinct morphogenic processes.<br />

In conclusion, this study identifies a prostasin-matriptase cell<br />

surface protease cascade whose activity must be suppressed by<br />

PLOS Genetics | www.plosgenetics.org 14 August 2012 | Volume 8 | Issue 8 | e1002937


Embryonic Cell Surface Serine Protease Cascade<br />

HAI-1 and HAI-2 to enable early embryonic ectoderm formation,<br />

placental morphogenesis, and neural tube closure.<br />

Materials and Methods<br />

Mouse strains<br />

All experiments were performed in an Association for Assessment<br />

and Accreditation of Laboratory Animal Care Internationalaccredited<br />

vivarium following Standard Operating Procedures. The<br />

studies were approved by the NIDCR <strong>Institut</strong>ional Animal Care<br />

and Use Committee. All studies were littermate controlled.<br />

Spint1 2/2 , Spint2 2/2 , St14 2/2 , Hgfr 2/2 , F2r 2/2 , F2rl1 2/2 ,<br />

Scnn1a 2/2 , and Prss8 fr/fr mice have been described<br />

[38,40,43,49,66,67,68]. Prostasin-deficient (Prss8 2/2 ) mice were<br />

generated by standard blastocyst injection of C57BL/6J-derived<br />

embryonic stem cells carrying a gene trap insertion in the Prss8 gene<br />

(clone IST10122F12, Texas A&M <strong>Institut</strong>e for Genomic Research,<br />

College Station, TX).<br />

Extraction of embryonic and perinatal tissues<br />

Breeding females were checked for vaginal plugs in the morning<br />

and the day on which the plug was found was defined as the first<br />

day of pregnancy (E0.5). Pregnant females were euthanized in the<br />

mid-day at designated time points by CO 2 asphyxiation. Embryos<br />

were extracted by Caesarian section and the individual embryos<br />

and placentae were dissected and processed. Visceral yolk sacs of<br />

individual embryos were washed twice in phosphate buffered<br />

saline, subjected to genomic DNA extraction and genotyped by<br />

PCR (see Table S1 for primer sequences). Newborn pups were<br />

euthanized by CO 2 inhalation at 0uC. For histological analysis, the<br />

embryos and newborn pups were fixed for 18–20 hrs in 4%<br />

paraformaldehyde (PFA) in PBS, processed into paraffin, sectioned,<br />

and stained with hematoxylin and eosin (H&E), or used for<br />

immunohistochemistry as described below. For histomorphometric<br />

analysis of placental labyrinth, the midline cross sections of<br />

plancetal tissues were stained with H&E and the thickness of the<br />

labyrinth was determined as the maximum perpendicular distance<br />

of fetal vessel from the chorionic trophoblast layer.<br />

Immunohistochemistry<br />

Antigens from 5 mm paraffin sections were retrieved by<br />

incubation for 10 min at 37uC with 10 mg/ml proteinase K<br />

(Fermentas, Hanover, MD) for HAI-1 staining, or by incubation<br />

for 20 min at 100uC in 0.01 M sodium citrate buffer, pH 6.0, for<br />

all other antigens. The sections were blocked with 2% bovine<br />

serum albumin in PBS, and incubated overnight at 4uC with<br />

rabbit anti-human CD31 (1:100, Santa Cruz Biotechnology, Santa<br />

Cruz, CA), goat anti-mouse HAI-1 (1:200, R&D Systems,<br />

Minneapolis, MN), mouse anti-human prostasin (1:200, BD<br />

Transduction Laboratories, San Jose, CA), sheep anti-human<br />

matriptase (1:200, R&D Systems) or ENaCc subunit (1:100,<br />

Sigma-Aldrich, St. Louis, MO) primary antibodies. Bound<br />

antibodies were visualized using biotin-conjugated anti-mouse, -<br />

rabbit, -sheep or -goat secondary antibodies (all 1:400, Vector<br />

Laboratories, Burlingame, CA) and a Vectastain ABC kit (Vector<br />

Laboratories) using 3,39-diaminobenzidine as the substrate (Sigma-Aldrich).<br />

All microscopic images were acquired on an<br />

Olympus BX40 microscope using an Olympus DP70 digital<br />

camera system (Olympus, Melville, NY).<br />

Protein extraction from mouse tissues<br />

Placentae were extracted from embryos at E10.5 or E11.5. The<br />

embryonic portion of each placenta was manually separated from<br />

maternal decidua using a dissecting microscope. The tissues were<br />

then homogenized in ice-cold 50 mM Tris/HCl, pH 8.0; 1% NP-<br />

40; 500 mM NaCl buffer and incubated on ice for 10 minutes.<br />

The lysates were centrifuged at 20,000 g for 10 min at 4uC to<br />

remove the tissue debris and the supernatant was used for further<br />

analysis as described below.<br />

Detection of active matriptase and prostasin in mouse<br />

embryonic tissues<br />

Lysates from two placentae of the same genotype were<br />

combined and pre-incubated with 100 ul GammaBind G<br />

Sepharose beads (GE Healthcare Bio-Sciences, Uppsala, Sweden)<br />

for 30 minutes at 4uC with gentle agitation. The samples were<br />

spun at 5,000 g for 1 min to remove the beads, and the<br />

supernatant was then incubated with 5 mg goat anti-mouse HAI-<br />

1 antibody (R&D Systems) and 100 ul of GammaBind G<br />

Sepharose beads for 3 hours at 4uC. The samples were spun at<br />

5,000 g for 1 min, the supernatant was removed, and the beads<br />

were washed 3 times with 1 ml ice-cold 50 mM Tris/HCl,<br />

pH 8.0; 1% NP-40; 500 mM NaCl buffer. The beads were then<br />

mixed with 30 ul of 16SDS loading buffer (Invitrogen, Carlsbad,<br />

CA) with 0.25 M b-mercaptoethanol, incubated for 5 min at<br />

99uC, and cooled on ice for 2 minutes. The samples were spun at<br />

5,000 g for 1 min and the released proteins were resolved by SDS-<br />

PAGE (4–12% polyacrylamide gel) and analyzed by western blot<br />

using mouse anti-human prostasin (1:250, BD Transduction Labs)<br />

or sheep anti-human matriptase (1:500, R&D Systems) primary<br />

antibodies and goat anti-mouse (DakoCytomation) or donkey antisheep<br />

(Sigma-Aldrich) secondary antibodies (both 1:1000) conjugated<br />

to alkaline phosphatase, and visualized using nitro-blue<br />

tetrazolium and 5-bromo-4-chloro-39-indolyphosphate.<br />

Amiloride injections<br />

Five ug per g of body weight of amiloride (Sigma-Aldrich) in<br />

10% DMSO in PBS was administered to pregnant females by<br />

intraperitoneal injection every 24 hours starting on E5.5. Embryos<br />

were extracted on E9.5 by Caesarian section and genotyped as<br />

described, and scored for neural tube closure defects.<br />

Generation of soluble recombinant wildtype, catalytically<br />

inactive S238A, and V170D prostasin zymogens<br />

The generation of pIRES2-EGFP-prostasin has been described<br />

[26]. Substitution of the native prostasin activation site (APQAR)<br />

by the enteropeptidase-dependent cleavage site (DDDDK), and<br />

either the S238A or V170D point mutations were introduced<br />

using the QuickChange Kit (Stratagene, La Jolla, CA) and the<br />

following primers, respectively: 59-GCTCCCTGCGGTGTGG-<br />

CCCCCCAAGCACGCATCACAGGTGGCAGC-39, 59-GAC-<br />

GCCTGCCAGGGTGACGCTGGGGGCCCACTCTCCTGC-<br />

39, and 59-GGCCTCCACTGCACTGACACTGGCTGGGGT-<br />

CAT-39. Successful mutagenesis was verified by sequencing of<br />

both strands of the resulting cDNA. Expression plasmids carrying<br />

individual mutations were transiently transfected into HEK-293T<br />

cells using Turbofect (Fermentas). The cells were grown for two<br />

days and soluble recombinant prostasin was prepared by<br />

treatment of cells with phosphatidylinositol-specific phospholipase<br />

C (Sigma-Aldrich) as described previously [26].<br />

Determination of enzymatic activity of recombinant<br />

prostasin variants<br />

Recombinant wildtype, V170D Frizzy or catalytically inactive<br />

S238A prostasin zymogen variants were first incubated with 5.1 U<br />

recombinant bovine enteropeptidase (Novagen, Cambridge, MA)<br />

overnight at 37uC in enterokinase buffer (Novagen). Following<br />

PLOS Genetics | www.plosgenetics.org 15 August 2012 | Volume 8 | Issue 8 | e1002937


Embryonic Cell Surface Serine Protease Cascade<br />

enteropeptidase removal using the Enterokinase Removal Kit<br />

(Sigma-Aldrich), the protein concentration was estimated by<br />

western blot of serially diluted proteins using a reference with<br />

known protein concentration. For substrate hydrolysis assays, the<br />

activated prostasin variants (62.5 nM) were incubated with the<br />

fluorogenic substrate pERTKR-AMC (50 mM final concentration)<br />

(R&D systems) at 37uC in 50 mM NaCl, 50 mM Tris-HCl<br />

pH 8.8, 0.01% Tween-20 buffer, and the fluorescence was<br />

measured using a Wallac plate reader (Perkin Elmer, Waltham,<br />

MA). Each measurement was performed in triplicate. For serpin<br />

complex formation, prostasin variants were diluted in 50 mM<br />

Tris-HCl, pH 9.0, 50 mM NaCl, 0.01% Tween 20 to a final<br />

concentration of 150 nM, incubated with 250 ng recombinant<br />

human protease nexin-1 (PN-1) (R&D Systems) for 1 h at 37uC,<br />

and analyzed using 12% reducing SDS-PAGE and western<br />

blotting, using a monoclonal anti-prostasin antibody (BD Transduction<br />

Laboratories).<br />

Matriptase activation and SRE–luciferase assay<br />

HEK 293 cells were plated in 24-well plates and grown in<br />

DMEM supplemented with 10% FBS for 24 h. Cells were cotransfected<br />

with pSRE-firefly luciferase (50 ng), pRL-Renilla<br />

luciferase (20 ng), pcDNA 3.1 Par2 (100 ng) (Missouri S&T<br />

cDNA Resource Center) using Lipofectamine and Plus reagent<br />

(Invitrogen), pcDNA 3.1 expression vectors containing wildtype<br />

human matriptase or catalytically dead matriptase (S805A), full<br />

length human HAI-1 [69] and empty pcDNA 3.1 vector to<br />

equalize the total amount of transfected DNA. After 36 h the cells<br />

were serum starved over night and then stimulated with 100 nM<br />

recombinant human soluble prostasin (R&D) or vehicle for 6 h.<br />

Cell were lysed and luciferase activity was determined using the<br />

dual luciferase assay kit (Promega, Madison, WI) according to the<br />

manufacturer’s instructions. Chemiluminiscence was measured<br />

using Microtiter Plate Luminometer (Dynex Technologies,<br />

Chantilly, VA) and the SRE activation was determined as the<br />

ratio of firefly to Renilla luciferase counts. The assay was<br />

performed two times in duplicates.<br />

Supporting Information<br />

Figure S1 (A) Western blot detection of protein nexin-1 (PN-1).<br />

Wildtype zymogen (lanes 1 and 2), activated wildtype (lanes 3 and<br />

4), V170D (frizzy) zymogen (lanes 5 and 6), activated V170D<br />

(lanes 7 and 8), S238A zymogen (lanes 9 and 10), and activated<br />

S238A (lanes 11 and 12) prostasin variants were incubated with<br />

(lanes 2, 4, 6, 8, 10, and 12) or without (lanes 1, 3, 5, 7, 9, and 11)<br />

250 ng of recombinant human PN-1. Position of PN-1, and<br />

predicted position of prostasin/PN-1 complexes (not detected by<br />

anti-PN-1 antibody presumably due to significant molecular<br />

rearrangement of PN-1 in the complex with the protease) are<br />

indicated. Positions of molecular weight markers (kDa) are shown<br />

on left. (B) Western blot detection of active matriptase in the fetal<br />

part of the E11.5 placentas of one matriptase-deficient<br />

(St14 2/2 ;Prss8 +/+ ) (lane 1), three wildtype (Prss8 +/+ and St14 +/+ )<br />

(lanes 2,3, and 4), and three prostasin-deficient (St14 +/+ ;Prss8 2/2 )<br />

(lanes 5, 6, and 7) embryos after anti-HAI-1 immunoprecipitation.<br />

A 30 kDa band representing the active serine protease domain of<br />

matriptase (Mat SPD) was present in extracts from wildtype, but<br />

not in matriptase- or prostasin-deficient placentas. (C) Distribution<br />

of Spint2 genotypes at E9.5 in offspring from interbred Spint2 +/2<br />

breeding pairs treated with the ENaC inhibitor, amiloride, at<br />

E5.5–8.5. No Spint2 2/2 embryos were observed.<br />

(TIF)<br />

Table S1 Sequences of PCR primers used for mouse genotyping.<br />

(DOCX)<br />

Acknowledgments<br />

We thank Dr. Mary Jo Danton for critically reviewing this manuscript.<br />

Histology was performed by Histoserv, Germantown, Maryland, United<br />

States of America.<br />

Author Contributions<br />

Conceived and designed the experiments: RS THB JSG. Performed the<br />

experiments: RS KUS PK NAS SG EC KKH SF. Analyzed the data: RS<br />

KUS PK NAS EC THB KKH. Contributed reagents/materials/analysis<br />

tools: SG LKV EH. Wrote the paper: RS THB.<br />

References<br />

1. Szabo R, Bugge TH (2011) Membrane anchored serine proteases in cell and<br />

developmental biology. Annu Rev Cell and Developmental Biology 27: 213–235.<br />

2. Bugge TH, Antalis TM, Wu Q (2009) Type II transmembrane serine proteases.<br />

J Biol Chem 284: 23177–23181.<br />

3. Antalis TM, Buzza MS, Hodge KM, Hooper JD, Netzel-Arnett S (2010) The<br />

cutting edge: membrane-anchored serine protease activities in the pericellular<br />

microenvironment. Biochem J 428: 325–346.<br />

4. Guipponi M, Tan J, Cannon PZ, Donley L, Crewther P, et al. (2007) Mice<br />

deficient for the type II transmembrane serine protease, TMPRSS1/hepsin,<br />

exhibit profound hearing loss. Am J Pathol 171: 608–616.<br />

5. Scott HS, Kudoh J, Wattenhofer M, Shibuya K, Berry A, et al. (2001) Insertion of<br />

beta-satellite repeats identifies a transmembrane protease causing both congenital<br />

and childhood onset autosomal recessive deafness. Nat Genet 27: 59–63.<br />

6. Fasquelle L, Scott HS, Lenoir M, Wang J, Rebillard G, et al. (2010) Tmprss3, a<br />

transmembrane serine protease deficient in human DFNB8/10 deafness, is<br />

critical for Cochlear hair cell survival at the onset of hearing. J Biol Chem 286:<br />

17383–17397.<br />

7. List K, Szabo R, Wertz PW, Segre J, Haudenschild CC, et al. (2003) Loss of<br />

proteolytically processed filaggrin caused by epidermal deletion of Matriptase/<br />

MT-SP1. J Cell Biol 163: 901–910.<br />

8. Leyvraz C, Charles RP, Rubera I, Guitard M, Rotman S, et al. (2005) The<br />

epidermal barrier function is dependent on the serine protease CAP1/Prss8.<br />

J Cell Biol 170: 487–496.<br />

9. Buzza MS, Netzel-Arnett S, Shea-Donohue T, Zhao A, Lin CY, et al. (2010)<br />

Membrane-anchored serine protease matriptase regulates epithelial barrier<br />

formation and permeability in the intestine. Proc Natl Acad Sci U S A 107:<br />

4200–4205.<br />

10. List K, Kosa P, Szabo R, Bey AL, Wang CB, et al. (2009) Epithelial integrity is<br />

maintained by a matriptase-dependent proteolytic pathway. Am J Pathol 175:<br />

1453–1463.<br />

11. Zheng XL, Kitamoto Y, Sadler JE (2009) Enteropeptidase, a type II<br />

transmembrane serine protease. Front Biosci (Elite Ed) 1: 242–249.<br />

12. Dries DL, Victor RG, Rame JE, Cooper RS, Wu X, et al. (2005) Corin gene<br />

minor allele defined by 2 missense mutations is common in blacks and associated<br />

with high blood pressure and hypertension. Circulation 112: 2403–2410.<br />

13. Wang W, Liao X, Fukuda K, Knappe S, Wu F, et al. (2008) Corin variant<br />

associated with hypertension and cardiac hypertrophy exhibits impaired<br />

zymogen activation and natriuretic peptide processing activity. Circ Res 103:<br />

502–508.<br />

14. Yan W, Wu F, Morser J, Wu Q (2000) Corin, a transmembrane cardiac serine<br />

protease, acts as a pro-atrial natriuretic peptide-converting enzyme. Proc Natl<br />

Acad Sci U S A 97: 8525–8529.<br />

15. Du X, She E, Gelbart T, Truksa J, Lee P, et al. (2008) The serine protease<br />

TMPRSS6 is required to sense iron deficiency. Science 320: 1088–1092.<br />

16. Finberg KE, Heeney MM, Campagna DR, Aydinok Y, Pearson HA, et al.<br />

(2008) Mutations in TMPRSS6 cause iron-refractory iron deficiency anemia<br />

(IRIDA). Nat Genet 40: 569–571.<br />

17. Netzel-Arnett S, Bugge TH, Hess RA, Carnes K, Stringer BW, et al. (2009) The<br />

glycosylphosphatidylinositol-anchored serine protease PRSS21 (testisin) imparts<br />

murine epididymal sperm cell maturation and fertilizing ability. Biol Reprod 81:<br />

921–932.<br />

18. Kawano N, Kang W, Yamashita M, Koga Y, Yamazaki T, et al. (2010) Mice<br />

lacking two sperm serine proteases, ACR and PRSS21, are subfertile, but the<br />

mutant sperm are infertile in vitro. Biol Reprod 83: 359–369.<br />

PLOS Genetics | www.plosgenetics.org 16 August 2012 | Volume 8 | Issue 8 | e1002937


Embryonic Cell Surface Serine Protease Cascade<br />

19. Heinz-Erian P, Muller T, Krabichler B, Schranz M, Becker C, et al. (2009)<br />

Mutations in SPINT2 Cause a Syndromic Form of Congenital Sodium<br />

Diarrhea. Am J Hum Genet, 84: 188–196.<br />

20. Szabo R, Kosa P, List K, Bugge TH (2009) Loss of matriptase suppression<br />

underlies spint1 mutation-associated ichthyosis and postnatal lethality.<br />

Am J Pathol 174: 2015–2022.<br />

21. Nagaike K, Kawaguchi M, Takeda N, Fukoshima T, Sawaguchi A, et al. (2008)<br />

Defect of hepatocyte growth factor activator inhibitor type 1/serine protease<br />

inhibior, Kunitz type 1 (Hai-1/Spint1) leads to ichthyosis-like condition and<br />

abnormal hair development in mice. American Journal of Pathology 173: 1–12.<br />

22. Kawaguchi M, Takeda N, Hoshiko S, Yorita K, Baba T, et al. (2011)<br />

Membrane-bound serine protease inhibitor HAI-1 is required for maintenance<br />

of intestinal epithelial integrity. Am J Pathol 179: 1815–1826.<br />

23. List K, Currie B, Scharschmidt TC, Szabo R, Shireman J, et al. (2007)<br />

Autosomal ichthyosis with hypotrichosis syndrome displays low matriptase<br />

proteolytic activity and is phenocopied in ST14 hypomorphic mice. J Biol Chem<br />

282: 36714–36723.<br />

24. Alef T, Torres S, Hausser I, Metze D, Tursen U, et al. (2009) Ichthyosis,<br />

Follicular Atrophoderma, and Hypotrichosis Caused by Mutations in ST14 Is<br />

Associated with Impaired Profilaggrin Processing. J Invest Dermatol 129: 862–<br />

869.<br />

25. Chen YW, Wang JK, Chou FP, Chen CY, Rorke EA, et al. (2010) Regulation of<br />

the matriptase-prostasin cell surface proteolytic cascade by hepatocyte growth<br />

factor activator inhibitor-1 during epidermal differentiation. J Biol Chem 285:<br />

31755–31762.<br />

26. Netzel-Arnett S, Currie BM, Szabo R, Lin CY, Chen LM, et al. (2006) Evidence<br />

for a matriptase-prostasin proteolytic cascade regulating terminal epidermal<br />

differentiation. J Biol Chem 281: 32941–32945.<br />

27. Kim C, Lee HS, Lee D, Lee SD, Cho EG, et al. (2011) Epithin/PRSS14<br />

proteolytically regulates angiopoietin receptor Tie2 during transendothelial<br />

migration. Blood 117: 1415–1424.<br />

28. Ustach CV, Huang W, Conley-LaComb MK, Lin CY, Che M, et al. (2010) A<br />

novel signaling axis of matriptase/PDGF-D/ss-PDGFR in human prostate<br />

cancer. Cancer Res 70: 9631–9640.<br />

29. Lee SL, Dickson RB, Lin CY (2000) Activation of hepatocyte growth factor and<br />

urokinase/plasminogen activator by matriptase, an epithelial membrane serine<br />

protease. J Biol Chem 275: 36720–36725.<br />

30. Owen KA, Qiu D, Alves J, Schumacher AM, Kilpatrick LM, et al. (2010)<br />

Pericellular activation of hepatocyte growth factor by the transmembrane serine<br />

proteases matriptase and hepsin, but not by the membrane-associated protease<br />

uPA. Biochem J 426: 219–228.<br />

31. Takeuchi T, Harris JL, Huang W, Yan KW, Coughlin SR, et al. (2000) Cellular<br />

localization of membrane-type serine protease 1 and identification of proteaseactivated<br />

receptor-2 and single-chain urokinase-type plasminogen activator as<br />

substrates. J Biol Chem 275: 26333–26342.<br />

32. Bhatt AS, Welm A, Farady CJ, Vasquez M, Wilson K, et al. (2007) Coordinate<br />

expression and functional profiling identify an extracellular proteolytic signaling<br />

pathway. Proc Natl Acad Sci USA 104: 5771–5776.<br />

33. Bhatt AS, Erdjument-Bromage H, Tempst P, Craik CS, Moasser MM (2005)<br />

Adhesion signaling by a novel mitotic substrate of src kinases. Oncogene 24:<br />

5333–5343.<br />

34. Vuagniaux G, Vallet V, Jaeger NF, Hummler E, Rossier BC (2002) Synergistic<br />

activation of ENaC by three membrane-bound channel-activating serine<br />

proteases (mCAP1, mCAP2, and mCAP3) and serum- and glucocorticoidregulated<br />

kinase (Sgk1) in Xenopus Oocytes. J Gen Physiol 120: 191–201.<br />

35. Andreasen D, Vuagniaux G, Fowler-Jaeger N, Hummler E, Rossier BC (2006)<br />

Activation of epithelial sodium channels by mouse channel activating proteases<br />

(mCAP) expressed in Xenopus oocytes requires catalytic activity of mCAP3 and<br />

mCAP2 but not mCAP1. J Am Soc Nephrol 17: 968–976.<br />

36. Jin X, Yagi M, Akiyama N, Hirosaki T, Higashi S, et al. (2006) Matriptase<br />

activates stromelysin (MMP-3) and promotes tumor growth and angiogenesis.<br />

Cancer Sci 97: 1327–1334.<br />

37. Kilpatrick LM, Harris RL, Owen KA, Bass R, Ghorayeb C, et al. (2006)<br />

Initiation of plasminogen activation on the surface of monocytes expressing the<br />

type II transmembrane serine protease matriptase. Blood 108: 2616–2623.<br />

38. List K, Haudenschild CC, Szabo R, Chen W, Wahl SM, et al. (2002)<br />

Matriptase/MT-SP1 is required for postnatal survival, epidermal barrier<br />

function, hair follicle development, and thymic homeostasis. Oncogene 21:<br />

3765–3779.<br />

39. List K, Szabo R, Molinolo A, Nielsen BS, Bugge TH (2006) Delineation of<br />

matriptase protein expression by enzymatic gene trapping suggests diverging<br />

roles in barrier function, hair formation, and squamous cell carcinogenesis.<br />

Am J Pathol 168: 1513–1525.<br />

40. Szabo R, Molinolo A, List K, Bugge TH (2007) Matriptase inhibition by<br />

hepatocyte growth factor activator inhibitor-1 is essential for placental<br />

development. Oncogene 26: 1546–1556.<br />

41. Fan B, Brennan J, Grant D, Peale F, Rangell L, et al. (2007) Hepatocyte growth<br />

factor activator inhibitor-1 (HAI-1) is essential for the integrity of basement<br />

membranes in the developing placental labyrinth. Dev Biol 303: 222–230.<br />

42. Tanaka H, Nagaike K, Takeda N, Itoh H, Kohama K, et al. (2005) Hepatocyte<br />

Growth Factor Activator Inhibitor Type 1 (HAI-1) Is Required for Branching<br />

Morphogenesis in the Chorioallantoic Placenta. Mol Cell Biol 25: 5687–5698.<br />

43. Szabo R, Hobson JP, Christoph K, Kosa P, List K, et al. (2009) Regulation of<br />

cell surface protease matriptase by HAI2 is essential for placental development,<br />

neural tube closure and embryonic survival in mice. Development 136: 2653–<br />

2663.<br />

44. Mitchell KJ, Pinson KI, Kelly OG, Brennan J, Zupicich J, et al. (2001)<br />

Functional analysis of secreted and transmembrane proteins critical to mouse<br />

development. Nat Genet 28: 241–249.<br />

45. Camerer E, Barker A, Duong DN, Ganesan R, Kataoka H, et al. (2010) Local<br />

protease signaling contributes to neural tube closure in the mouse embryo. Dev<br />

Cell 18: 25–38.<br />

46. Szabo R, Rasmussen AL, Moyer AB, Kosa P, Schafer J, et al. (2011) c-Metinduced<br />

epithelial carcinogenesis is initiated by the serine protease matriptase.<br />

Oncogene 30: 2003–2016.<br />

47. Bladt F, Riethmacher D, Isenmann S, Aguzzi A, Birchmeier C (1995) Essential<br />

role for the c-met receptor in the migration of myogenic precursor cells into the<br />

limb bud. Nature 376: 768–771.<br />

48. Uehara Y, Minowa O, Mori C, Shiota K, Kuno J, et al. (1995) Placental defect<br />

and embryonic lethality in mice lacking hepatocyte growth factor/scatter factor.<br />

Nature 373: 702–705.<br />

49. Spacek DV, Perez AF, Ferranti KM, Wu LK, Moy DM, et al. (2010) The mouse<br />

frizzy (fr) and rat ‘hairless’ (frCR) mutations are natural variants of protease<br />

serine S1 family member 8 (Prss8). Exp Dermatol 19: 527–532.<br />

50. Hedstrom L (2002) Serine protease mechanism and specificity. Chem Rev 102:<br />

4501–4524.<br />

51. Rossier BC, Stutts MJ (2009) Activation of the epithelial sodium channel (ENaC)<br />

by serine proteases. Annu Rev Physiol 71: 361–379.<br />

52. Planes C, Randrianarison NH, Charles R-P, Frateschi S, Cluzeeaud D, et al.<br />

(2009) ENaC-mediated alveolar fluid clearance and lung fluid balance depend<br />

on the channel-activating protease 1. EMBO Molecular Medicine 2: 26–37.<br />

53. Miller TA, Scott WJ, Jr. (1992) Abnormalities in ureter and kidney development<br />

in mice given acetazolamide-amiloride or dimethadione (DMO) during<br />

embryogenesis. Teratology 46: 541–550.<br />

54. Canessa CM, Schild L, Buell G, Thorens B, Gautschi I, et al. (1994) Amiloridesensitive<br />

epithelial Na+ channel is made of three homologous subunits. Nature<br />

367: 463–467.<br />

55. Hummler E, Barker P, Gatzy J, Beermann F, Verdumo C, et al. (1996) Early<br />

death due to defective neonatal lung liquid clearance in alpha-ENaC-deficient<br />

mice. Nat Genet 12: 325–328.<br />

56. Oberst MD, Williams CA, Dickson RB, Johnson MD, Lin CY (2003) The<br />

activation of matriptase requires its noncatalytic domains, serine protease<br />

domain, and its cognate inhibitor. J Biol Chem 278: 26773–26779.<br />

57. Shipway A, Danahay H, Williams JA, Tully DC, Backes BJ, et al. (2004)<br />

Biochemical characterization of prostasin, a channel activating protease.<br />

Biochem Biophys Res Commun 324: 953–963.<br />

58. Denda K, Shimomura T, Kawaguchi T, Miyazawa K, Kitamura N (2002)<br />

Functional characterization of Kunitz domains in hepatocyte growth factor<br />

activator inhibitor type 1. J Biol Chem 277: 14053–14059.<br />

59. Kirchhofer D, Peek M, Lipari MT, Billeci K, Fan B, et al. (2005) Hepsin<br />

activates pro-hepatocyte growth factor and is inhibited by hepatocyte growth<br />

factor activator inhibitor-1B (HAI-1B) and HAI-2. FEBS Lett 579: 1945–1950.<br />

60. Kirchhofer D, Peek M, Li W, Stamos J, Eigenbrot C, et al. (2003) Tissue<br />

expression, protease specificity, and Kunitz domain functions of hepatocyte<br />

growth factor activator inhibitor-1B (HAI-1B), a new splice variant of HAI-1.<br />

J Biol Chem 278: 36341–36349.<br />

61. Herter S, Piper DE, Aaron W, Gabriele T, Cutler G, et al. (2005) Hepatocyte<br />

growth factor is a preferred in vitro substrate for human hepsin, a membraneanchored<br />

serine protease implicated in prostate and ovarian cancers. Biochem J<br />

390: 125–136.<br />

62. Fan B, Wu TD, Li W, Kirchhofer D (2005) Identification of hepatocyte growth<br />

factor activator inhibitor-1B as a potential physiological inhibitor of prostasin.<br />

J Biol Chem 280: 34513–34520.<br />

63. Delaria KA, Muller DK, Marlor CW, Brown JE, Das RC, et al. (1997)<br />

Characterization of placental bikunin, a novel human serine protease inhibitor.<br />

J Biol Chem 272: 12209–12214.<br />

64. Qin L, Denda K, Shimomura T, Kawaguchi T, Kitamura N (1998) Functional<br />

characterization of Kunitz domains in hepatocyte growth factor activator<br />

inhibitor type 2. FEBS Lett 436: 111–114.<br />

65. Copp AJ, Greene ND (2010) Defining a PARticular pathway of neural tube<br />

closure. Dev Cell 18: 1–2.<br />

66. Hummler E, Merillat AM, Rubera I, Rossier BC, Beermann F (2002)<br />

Conditional gene targeting of the Scnn1a (alphaENaC) gene locus. Genesis<br />

32: 169–172.<br />

67. Connolly AJ, Ishihara H, Kahn ML, Farese RV, Jr., Coughlin SR (1996) Role of<br />

the thrombin receptor in development and evidence for a second receptor.<br />

Nature 381: 516–519.<br />

68. Lindner JR, Kahn ML, Coughlin SR, Sambrano GR, Schauble E, et al. (2000)<br />

Delayed onset of inflammation in protease-activated receptor-2- deficient mice.<br />

J Immunol 165: 6504–6510.<br />

69. Oberst MD, Chen LL, Kiyomiya K-I, Williams CA, Lee M-S, et al. (2005).<br />

HAI-1 regulates activation and expression of matriptase, a membrane-bound<br />

serine protease. Am J Physiol Cell Physiol 289: C462–C470.<br />

PLOS Genetics | www.plosgenetics.org 17 August 2012 | Volume 8 | Issue 8 | e1002937


Paper III<br />

Novel assay for detection of active matriptase<br />

Sine Godiksen 1,2 , and Lotte K. Vogel 2<br />

More authors to be added<br />

1 Department of Biology, University of Copenhagen, Copenhagen, Denmark<br />

2 Department of Cellular and Molecular Medicine, University of Copenhagen, Copenhagen,<br />

Denmark<br />

Manuscript in progress<br />

65


Novel assay for detection of active matriptase<br />

Sine Godiksen 1,2 and Lotte K. Vogel 2<br />

Other authors to be added<br />

1 Department of Biology, University of Copenhagen, Copenhagen, Denmark<br />

2 Department of Cellular and Molecular Medicine, University of Copenhagen, Copenhagen,<br />

Denmark<br />

Matriptase is a member of the family of type II transmembrane serine proteases that is crucial<br />

for development and maintenance of several epithelial tissues. Matriptase is synthesized as a<br />

zymogen that is converted by activation site cleavage to a disulfide-linked active form. This<br />

occurs by autoactivation, which has led to the hypothesis that matriptase functions at the<br />

pinnacle of several protease induced signal cascades involved in epithelial tissue formation and<br />

maintenance. Matriptase can be detected in either its zymogen form or in a complex with its<br />

cognate inhibitor hepatocyte growth factor activator inhibitor 1 (HAI-1) in cells, whereas the<br />

active form has been difficult to detect. In this study, we have established an assay to detect<br />

active matriptase using a peptide substrate-based chloromethyl ketone (Cmk) inhibitor. Our<br />

assay is based on the assumption that matriptase able to react with a Cmk peptide inhibitor is<br />

equivalent to active matriptase. In order to detect covalently Cmk peptide-bound matriptase,<br />

we designed a Cmk peptide inhibitor with a biotin moiety in the N-terminal of the small peptide<br />

that allows for streptavidin pull-down and subsequent analysis by Western blotting. This study<br />

presents a novel assay for detection active matriptase in living cells. The assay can be easily<br />

applied in other cell systems and other species.<br />

Abbreviations used: Cmk, chloromethyl ketone; HAI-1, hepatocyte growth factor activator<br />

inhibitor-1; RT, room temperature; SPD: serine protease domain<br />

Introduction<br />

Matriptase (also known as MT-SP1, epithin, TADG-15 and SNC19) is a member of the matriptase<br />

subfamily of type II transmembrane serine proteases. This protease is expressed in most epithelial<br />

cells and has pleiotropic roles in mouse epithelial homeostasis [1-5]. In humans, mutations in the<br />

ST14 gene expressing matriptase, is the underlying cause of congenital ichthyosis [6-8]. More<br />

recently, matriptase was identified as an initiator of the runaway kallikrein protease cascade<br />

leading to an autosomal recessive form of ichthyosis referred to as Netherton syndrome [9].


Ablation of the ST14 gene in mice shows that the protease is essential for postnatal survival.<br />

Matriptase knock-out mice display defects in epidermal barrier function, hair follicle<br />

development, pro-filaggrin processing, and in thymic homeostasis and die within 48 hrs of birth<br />

[5;10]. Tissue-specific or postnatal ablation of matriptase in mice cause severe organ dysfunction,<br />

generalized epithelial demise, and demonstrate a principal and global function of matriptase in<br />

promoting the formation of paracellular permeability barriers in simple and stratified epithelia [4].<br />

Important knowledge about matriptase has also been derived from ablation of the physiological<br />

inhibitors of matriptase; hepatocyte growth factor activator inhibitor-1 (HAI-1) and -2 (HAI-2).<br />

Genetic inactivation of either HAI-1 or HAI-2 leads to failure of placental labyrinth formation<br />

[11;12]. This defect can be completely rescued by simultaneously reducing or eliminating<br />

matriptase expression [12;13]. HAI-1 deficiency in both zebrafish and chimeric mice leads to fatal<br />

defects in epidermal integrity [14-17]. Simultaneous ablation of matriptase in both model systems<br />

completely eliminates these defects [14;16]. Additionally, ablation of HAI-2 in mice leads to<br />

defects in neural tube closure that are partially rescued by genetic inactivation of matriptase<br />

[12;18].<br />

Deregulated matriptase can promote carcinogenesis, as a modest overexpression of wild-type<br />

(WT) matriptase in the epidermis of transgenic mice is sufficient to induce spontaneous squamous<br />

cell carcinoma formation and to strongly potentiate chemical-induced skin carcinogenesis. The<br />

oncogenic effect is mediated through activation of the c-Met-Akt-mTor pathway by matriptasedependent<br />

activation of the cMet ligand; pro-HGF/SF [5;19;20]. A simultaneous increase in<br />

expression of HAI-1 completely negates the oncogenic effect of matriptase overexpression [19].<br />

In colitis-associated colorectal cancer, matriptase was recently assigned as a critical tumorsuppressor<br />

gene. Selective genetic inactivation of matriptase in the intestinal epithelium lead to a<br />

compromised intestinal barrier function associated with chronic inflammation and finally<br />

formation of colon adenocarcinomas in the mouse gastrointestinal tract [21].<br />

Matriptase is synthesized as an 855 amino acid single chain zymogen that undergoes two<br />

successive cleavages to become active. First, matriptase is cleaved in the sea urchin sperm<br />

protein, enteropeptidase, and agrin (SEA) domain by non-enzymatic hydrolysis and subsequently<br />

proteolytically processed at its canonical activation site motif after Arg614 in the serine protease<br />

domain. The serine protease domain (SPD) remains attached to the stem domain by a disulfide<br />

bridge. Shortly hereafter matriptase is inhibited by HAI-1 which leaves a short window of activity<br />

for the disulfide-linked active form of matriptase [22-24]. Using surface biotinylation, we have<br />

previously shown that matriptase undergoes activation site cleavage on the plasma membrane,<br />

and shortly after matriptase is endocytosed in complex with HAI-1 [25].<br />

Despite our knowledge about the important roles played by matriptase, both under normal<br />

physiological conditions as well as in a range of pathological conditions, our understanding of<br />

matriptase and the mechanisms regulating its proteolytic activity is poor. Many epithelial cells<br />

express matriptase that can be detected either in the zymogen or the HAI-1-complexed form<br />

[1;2;22;23;26]. However, the active non-inhibited form, which is the presumably biological active<br />

form of matriptase, has been difficult to detect.


In the present study, we have established an assay to detect active matriptase based on the<br />

assumption that active matriptase is equivalent to matriptase capable of binding an inhibitor. The<br />

assay employs a chloromethyl ketone (Cmk) peptide inhibitor which combined with Western<br />

blotting specifically detects active matriptase.<br />

Materials and methods<br />

Chromogenic assay<br />

0.2 μM matriptase SPD was prepared with 20 mM HEPES pH 7.4, 140 mM NaCl supplemented<br />

with 0.1% BSA (Sigma-Aldrich) in wells of a 96 well plate. Stocks of varying concentrations (5 nM<br />

or 50 μM) of biotin-RQRR-Cmk peptide (American Peptide) were incubated in the same buffer at<br />

37°C for up to 3 hours. At specific time points 100 μl of inhibitor solution was added to the well<br />

containing the indicated concentration of the active catalytic domain of matriptase [27]. Following<br />

additional 10 min incubation at 37°C 10 μl 6.3 mM chromogenic substrate H-D-Isoleucil-L-prolyl-Larginine-p-nitroaniline<br />

(cat. no. S2288, Chromogenix) was added and substrate conversion was<br />

followed in a standard plate reader by 30 min continuous measurements of the absorbance at<br />

405 nm at 37°C. The rate of substrate turnover was determined from the color development<br />

resulting from a pseudo first order reaction due to a substrate concentration far greater than the<br />

expected nM range of protease. To test pH effects, HEPES was replaced by 20 mM citric acid<br />

buffer pH 6.0.<br />

Cell culture<br />

Caco-2 cells were grown in minimal essential medium supplemented with 2 mM L-glutamine, 20%<br />

fetal bovine serum (Gibco), 1 x non essential amino acids, 100 units/ml penicillin and 100 μg/ml<br />

streptomycin (Invitrogen) at 37°C in an atmosphere of 5% CO 2 . For all experiments, 1-2 × 10 6 cells<br />

were seeded into 35 mm tissue culture plates or 0.4 μm-pore-size 24 mm Transwell® filters<br />

(Corning) allowing separate access to the apical and the basolateral plasma membrane. The cell<br />

culture medium was changed every day. Cells were grown until day 11 days post confluence, as<br />

indicated by the tightness of the cell monolayer for filter grown cells, before they were used in<br />

experiments. The tightness of filter-grown cells was assayed by filling the inner chamber to the<br />

brim and allowing it to equilibrate overnight.<br />

Isolation and short-term culture of primary keratinocytes from newborn mice<br />

Epidermis was isolated from newborn mice (d1-2) and grown in culture as described in [28]. In<br />

short, newborn pups were euthanized by decapitation and the torso was submerged in betadine<br />

and ethanol to sterilize the skin. The skin was incubated in 0.25% trypsin w/o EDTA (Sigma-


Aldrich) o/N at 4°C. The dermal portion was discarded and epidermis was minced to release<br />

keratinocytes. The epidermis was resuspended in 45 μM Ca 2+ /10%FBS/Keratinocyte-SFM<br />

(Invitrogen) media and was filtered through a 100 μm cell strainer and centrifuged to remove<br />

stratum corneum pieces. The cell pellet was resuspended in low calcium medium (45 μM<br />

Ca 2+ /Keratinocyte-SFM media) and plated on collagen (BD Biosciences) coated culture plates. Cell<br />

were grown in low calcium medium to sub-confluence and cell culture medium was changed<br />

every second day.<br />

Labeling with biotin-Arg-Gln-Arg-Arg-chloromethyl ketone (biotin-RQRR-Cmk) peptide inhibitor<br />

and S-NHS-SS-biotin<br />

Cells were washed twice; filter grown Caco-2 cells with PBS ++ (PBS supplemented with 0.7 mM<br />

CaCl 2 and 0.25 mM MgCl 2 ) and primary murine keratinocytes with PBS. For labeling of active<br />

matriptase, cells were incubated with 50 μM biotin-RQRR-Cmk (American Peptide) in serum-free<br />

MEM eagle with Earle´s supplemented with 0.2% NaHCO 3 , 100 units/ml penicillin and 100 μg/ml<br />

streptomycin (Invitrogen) at 37°C for the times indicated from the basolateral side. For acidinduced<br />

activation of matriptase, cells were labeled in a physiological phosphate buffer (25 mM<br />

Na 2 HPO 4 , 175 mM NaH 2 PO 4 ) pH 6 or pre-treated with physiological phosphate buffer pH 6 before<br />

labeling in serum-free MEM eagle with Earle´s supplemented with 0.2% NaHCO 3 , 100 units/ml<br />

penicillin and 100 μg/ml streptomycin. As a negative control, cells were labeled with 50 μM of a<br />

corresponding peptide without a CMK group; biotin-Arg-Gln-Arg-Arg (biotin-RQRR). For labeling of<br />

surface proteins, cells were biotinylated from the basolateral side with 1 mg/ml EZ-link Sulfo-<br />

NHS-SS-Biotin (Pierce) dissolved in PBS ++ for 30 min at 4°C. After peptide- and/or ordinary biotinlabeling,<br />

the cells were washed four times with ice-cold PBS ++ . In case of biotin-labeling, residual<br />

biotin was quenched with 50 mM glycine/PBS ++ for 5 min at 4°C and the cells were washed twice<br />

with PBS ++ . Cells were lysed in PBS containing 1% Triton X-100, 0.5% deoxycholate and protease<br />

inhibitors (10 mg/l benzamidine, 2 mg/l pepstatin A, 2 mg/l leupeptin, 2 mg/l antipain, and 2 mg/l<br />

chymostatin). Insoluble material was precipitated at 20,000×g for 20 min at 4°C and the<br />

supernatants were transferred to clean eppendorf tubes.<br />

Streptavidin pull down<br />

Cleared lysates were incubated for 2 hrs with end-over-end rotation at 4°C with 50 μl/24 mm<br />

filter pre-washed streptavidin-coated resin (Pierce), prepared as described by manufacturer. The<br />

streptavidin-coated resin was washed four times with 25 mM TRIS-HCl, 500 mM NaCl, 0.5% Triton<br />

X-100, pH 7.8, and three times with 10 mM TRIS-HCl, 150 mM NaCl, pH 7.8, and biotinylated<br />

proteins were eluted from the streptavidin-coated resin by boiling in SDS sample buffer.<br />

SDS-PAGE and Western blot<br />

Proteins were separated on 10% acrylamide gels and transferred to Immobilon-P PVDF<br />

membranes (Millipore). PVDF membranes were blocked with 10% non-fat dry milk in PBS<br />

containing 0.1% Tween-20 (PBST) for 1 hr at RT and were probed with primary antibody diluted in


1% non-fat dry milk in PBST at 4°C o/N. The next day the membranes were washed 3 times with<br />

PBST, followed by detection of bound primary antibody with horseradish peroxidase (HRP)-<br />

conjugated secondary antibody (Pierce) or alkaline phosphatase (AP)-conjugated secondary<br />

antibody (Sigma-Aldrich). After 3 washes with PBST the signal was developed using the ECL<br />

reagent Super Signal West Femto Maximum Sensitivity Substrate (Pierce) for HRP-conjugated<br />

secondary antibodies according to the protocol supplied by the manufacturer and visualized with<br />

a Fuji LAS1000 camera (Fujifilm Sweden AB) or by nitro-blue tetrazolium and 5- bromo-4-chloro-<br />

3´-indolyphosphate (Pierce) AP-conjugated secondary antibody.<br />

Antibodies<br />

The antibodies used for detection of matriptase in Western blotting were monoclonal mouse antihuman<br />

matriptase antibody M32 that reacts with the A chain recognizing total matriptase as a 70<br />

kDa band (both active and zymogen matriptase in boiled samples) and the 120-130 kDa complex<br />

of matriptase with HAI-1 (only in non-boiled samples) (3ug antibody/blot) [29]; monoclonal<br />

mouse anti-human matriptase antibody M69, which recognizes the 120-130kDa complex of<br />

matriptase with HAI-1 [29]; monoclonal mouse anti-human HAI-1 antibody M19, which recognizes<br />

free 55kDa HAI-1 and the 120-130kDa matriptase-HAI-1 complex [29]; polyclonal rabbit antihuman<br />

matriptase raised against the SPD of matriptase recognizing a 70 kDa band (both active<br />

and zymogen matriptase) under non-reducing conditions and the 70 kDa zymogen form and the<br />

30 kDa protease domain of cleaved matriptase under reducing conditions (Cat. no. IM1014,<br />

Calbiochem). For detection of matriptase in murine cells, sheep anti-matriptase was used (AF3946<br />

diluted 1:1000, R&D). Secondary antibodies include goat anti-mouse HRP-conjugated (10 ng/blot)<br />

(Pierce), goat anti-rabbit HRP-congugated (10 ng/blot) (Pierce) and donkey anti-sheep APcongugated<br />

(10 ng/blot) (Sigma-Aldrich).<br />

Results<br />

Biotin-RQRR-Cmk peptide inhibitor designed to react with active matriptase<br />

In order to detect active matriptase, we designed a peptide inhibitor biotin-RQRR-Cmk consisting<br />

of a tetra peptide; RQRR with an N-terminal biotin moiety and a C-terminal Cmk group. This<br />

inhibitor was designed based on a predicted substrate recognition sequence of matriptase [30].<br />

The Cmk group ensures that the protease-peptide interaction results in the formation of a<br />

covalent bond between the peptide and the protease by alkylation of the active site histidine<br />

attaching a biotin moiety to the active enzyme [31]. The biotin group allows for efficient<br />

concentration of active protease from a complex media by streptavidin precipitation [32].


Biotin-RQRR-Cmk inhibits the proteolytic activity of matriptase SPD in vitro<br />

To verify that biotin-RQRR-Cmk binds matriptase, we tested whether different concentrations of<br />

the peptide inhibitor were able to inhibit the proteolysis of a chromogenic substrate by purified<br />

recombinant matriptase serine protease domain (SPD). 50 μM biotin-RQRR-Cmk renders<br />

matriptase SPD unable to cleave the chromogenic substrate (Fig. 1A, black triangle). Although<br />

chloromethyl ketones are known to be hydrolysed in aqueous solutions with half lives in the order<br />

of 5-20 min [31], we found that an initial concentration of 50 μM biotin-RQRR-Cmk was still<br />

efficient after 180 min pre-incubation at 37°C and able to quench the activity of a sample<br />

containing 0.2 nM matriptase SPD (Fig. 1A, closed circle). A biotin-RQRR-Cmk concentration as low<br />

as 5nM is capable of inhibiting the peptidolytic activity of 0.2 nM matriptase SPD (fig. 1B, black<br />

triangle) even after 60 min of pre-incubation at 37°C (fig. 1B, open circle ), showing that biotin-<br />

RQRR-Cmk is very efficient near-stoichiometric inhibitor of matriptase SPD. Thus, biotin-RQRR-<br />

Cmk is a very efficient inhibitor of matriptase activity and relative stable in aqueous solutions.<br />

Biotin-RRQR-Cmk reacts with a subset of matriptase molecules on the surface of cultured cells<br />

Next, we wanted to test whether the biotin-RQRR-Cmk peptide inhibitor is able to bind active<br />

matriptase on the surface of cells in culture. Differentiated Caco-2 cells have an endogenous<br />

expression of matriptase that can be detected mainly as a zymogen form but also as a two-chain<br />

form in complex with HAI-1 [25;33]. We have previously shown that activation site cleavage of<br />

matriptase takes place on the basolateral plasma membrane of 11 days post-confluent Caco-2<br />

cells [25], indicating that these cells contains active matriptase on the basolateral plasma<br />

membrane.<br />

11 days post-confluent filter grown Caco-2 cells were treated with biotin-RQRR-Cmk from the<br />

basolateral side at 37°C for 2-180 min. Next, cells were lysed and biotin-RQRR-Cmk-labeled<br />

proteases were extracted from cleared lysates using streptavidin-coated resin. Proteins were<br />

released from the resin by boiling in SDS sample buffer. Due to substrate overlap of matriptase<br />

with other trypsin-like serine proteases [30], specificity of the assay is obtained by Western blot<br />

analysis using a matriptase specific antibody. To assess the steady state level of total surfaceassociated<br />

matriptase, parallel cultures were surface-biotinylated with S-NHS-SS-biotin at 4°C.<br />

Matriptase binds biotin-RQRR-Cmk as demonstrated by Western blotting (Fig. 2, lanes 4-6), and<br />

no matriptase could be detected when labeling with a control peptide; biotin-RQRR (Fig. 2, lane 2:<br />

CTRL). Biotin-RQRR-Cmk labeling displayed time dependence at 37°C as an increasing amount of<br />

matriptase was detected with increasing time of incubation with biotin-RQRR-Cmk, indicating that<br />

active matriptase is continuously generated. Comparison of steady state surface-associated<br />

matriptase by standard biotinylation techniques to the accumulated biotin-RQRR-Cmk labeled<br />

matriptase shows that only a fraction of surface-associated matriptase on Caco-2 cells is able to


ind biotin-RQRR-Cmk (compare lane 2 to lanes 3-6). Incubation with biotin-RQRR-Cmk for 180<br />

min at 4°C gave very low signals (data not shown) indicating that the steady state level of<br />

matriptase that is able to bind to biotin-RQRR-Cmk is low. Biotin-RQRR-Cmk also react with other<br />

serine proteases, as prostasin could be detected in Western blotting (data not shown),<br />

emphasizing that specificity of the assay depends on the antibody used for the Western blot<br />

analysis.<br />

Biotin-RRQR-Cmk does not react with the matriptase-HAI-1 complex<br />

After activation site cleavage, matriptase forms a reversible, but SDS-resistant complex with HAI-1<br />

that can be analyzed by SDS-PAGE under non-reducing conditions. In order to investigate whether<br />

biotin-RRQR-Cmk binds to the matriptase-HAI-1 complex, we took advantage of the fact that cells<br />

exposed to slightly acidic conditions has been reported to rapidly convert matriptase from the<br />

zymogen form into the matriptase-HAI-1 complex [24;26;34]. First, we tested whether exposure<br />

to slightly acidic conditions also in Caco-2 cells converts zymogen matriptase into a complex of<br />

activation site cleaved matriptase inhibited by HAI-1.<br />

11 days post-confluent filter grown Caco-2 cells were treated with physiological phosphate buffer<br />

pH 6.0 for 20 min (Fig. 3A, lanes 2, 4, 6, and 8) or left untreated (Fig. 3A, lanes 1, 3, 5, and 7). The<br />

two different lysates were investigated by Western blotting under non-boiled and non-reduced<br />

conditions (Fig. 3A, lanes 1-4 and 7-8) and under reducing conditions (Fig. 3A, lanes 5 and 6) using<br />

different antibodies. Untreated Caco-2 cells mainly contain 70 kDa zymogen matriptase (Fig. 3A,<br />

lanes 1 and 5) and to a minor degree the 130 kDa matriptase-HAI-1 complex (Fig. 3A, lane 1),<br />

whereas matriptase is detected as the 130 kDa matriptase-HAI-1 complex and to a minor degree<br />

70 kDa zymogen matriptase under non-boiling conditions after treatment with physiological<br />

phosphate buffer pH 6.0 (Fig. 3A, lane 2). The generation of the pH 6.0 induced 130 kDa<br />

matriptase-HAI-1 complex is confirmed by antibodies against the matriptase-HAI-1 complex (M69)<br />

(Fig. 3A, compare lanes 3 and 4) and a HAI-1 antibody that also recognizes the 130 kDa<br />

matriptase-HAI-1 complex (Fig. 3A, compare lanes 7 and 8). Analysis of the same samples under<br />

reducing conditions confirms the pH 6.0 induced activation site cleavage of matriptase, as<br />

matriptase was primarily detected as a band migrating at 30 kDa representing the released<br />

disulfide-linked SPD of matriptase after this treatment (Fig. 3A, lane 6). Thus, treatment with<br />

physiological phosphate buffer pH 6.0 induces activation site cleavage of matriptase and<br />

subsequent matriptase-HAI-1 complex formation in Caco-2 cells.<br />

Active matriptase has been shown to have pH optimum at pH 9 [30], we therefore tested the<br />

activity of recombinant matriptase SPD and the inhibitory capacity of biotin-RQRR-Cmk at pH 6 in<br />

the chromogenic assay previously described in Fig. 1. Matriptase SPD is able to cleave the<br />

chromogenic substrate at pH 6.0, although at a much lowered level, as compared to neutral pH<br />

(compare Fig. 3B, square to Fig. 1, square). When 50 μM biotin-RQRR-Cmk was added to the<br />

reaction, no cleavage of the substrate was observed (Fig. 3B, black triangle). Thus, matriptase SPD


is able to bind biotin-RQRR-Cmk at this lowered pH. This enabled us to test whether biotin-RRQR-<br />

Cmk binds to the matriptase-HAI-1 complex.<br />

To examine whether biotin-RQRR-Cmk reacts with the matriptase-HAI-1 complex, we employed<br />

physiological phosphate buffer pH 6.0 to induce matriptase activation. 11 days post-confluent<br />

filter grown Caco-2 cells were treated with 50 μM biotin-RQRR-Cmk from the basolateral side at<br />

37°C for 30 min under different conditions; labeling with biotin-RQRR-Cmk at pH 7.4 (Fig. 4, lanes<br />

3 and 7), labeling with biotin-RQRR-Cmk in physiological phosphate buffer pH 6.0 (Fig. 4, lanes 4<br />

and 8), or pre-treatment with physiological phosphate buffer pH 6.0 for 30 min, before labeling<br />

with biotin-RQRR-Cmk at pH 7.4 for 30 min (Fig. 4, lanes 5 and 9). Additional controls were lysates<br />

of untreated cells (Fig. 4, lane 1) and cells treated with physiological phosphate buffer pH 6.0 (Fig.<br />

4, lane 2) that show pH 6.0 induced matriptase activation and complex formation with HAI-1 (Fig.<br />

4, compare lane 1 and 2). In all cases, aliquots of the total lysates (Fig. 4, lanes 1-5) as well as the<br />

boiled streptavidin pull downs were analyzed in Western blot analysis (Fig. 4, lanes 6-9).<br />

Under the conditions used here, biotin-RQRR-Cmk is unable to dissociate the matriptase-HAIcomplex<br />

formed by pH 6 treatment (Fig. 4, compare lane 9 with lanes 4, 6 and 8). On the other<br />

hand, simultaneous treatment of Caco-2 cells with biotin-RQRR-Cmk and physiological phosphate<br />

buffer pH 6.0 hinders complex formation between matriptase and HAI-1 (Fig. 4, lanes 4 and 5) and<br />

instead more matriptase binds biotin-RQRR-Cmk as detected by Western blot analysis (Fig. 4,<br />

lanes 7 and 8). This observation is supporting the mutual sterical block of the active site in a<br />

competitive manner by either covalent binding to biotin-RQRR-Cmk or non-covalent, SDSresistant<br />

complex formation with HAI-1 [31;35]. Thus, HAI-1 and biotin-RQRR-Cmk competes for<br />

binding to active matriptase.<br />

Biotin-RRQR-Cmk reacts with zymogen matriptase<br />

CMK peptides have been widely used in studies of serine proteases and in most cases Cmkpeptides<br />

react only with the active form of the protease and not with the zymogen form except in<br />

a few cases of autoactivating proteases that have an intrinsic activity enabling the zymogen to<br />

reacts with the Cmk-peptide, one example is tPA [36;37]. Matriptase is also an autoactivating<br />

serine protease implying that the zymogen form has an intrinsic activity [26;38]. To investigate for<br />

an intrinsic activity of matriptase, we labeled 11 days post-confluent filter grown Caco-2 cells with<br />

biotin-RQRR-Cmk from the basolateral side for 180 min at 37°C. Biotin-RQRR-Cmk labeled<br />

proteases were precipitated with streptavidin-coated resin, and subjected to reducing SDS-PAGE.<br />

Samples were analyzed by Western blot analysis using an antibody against matriptase SPD. Under<br />

reducing conditions, matriptase could be detected both as the 70 kDa form (representing<br />

zymogen matriptase) and the 30 kDa form (representing activation site cleaved matriptase) with a<br />

stronger signal intensity of the 70 kDa form (Fig. 5, lane 2). No matriptase could be detected when<br />

labeling Caco-2 cells with a control peptide; biotin-RQRR (Fig. 5, lane 1: CTRL). Thus, both


activation site cleaved and zymogen matriptase is able to bind biotin-RQRR-Cmk indicating that<br />

zymogen matriptase has an intrinsic activity.<br />

Biotin-RRQR-Cmk also reacts with matriptase in other cell systems<br />

The described assay may easily be modified to detect active matriptase from other species as the<br />

specificity of the assay depends on the antibody used. To verify this, the presence of active<br />

matriptase was investigated in cultured primary murine keratinocytes. Keratinocytes from the<br />

skin of matriptase wild-type (WT) and matriptase deficient (KO) newborn pups were isolated,<br />

plated on collagen coated plastic and labeled with 50 μM biotin-RQRR-Cmk at subconfluence for<br />

180 min at 37°C or labeled with S-NHS-SS-biotin to assess surface-associated matriptase. As a<br />

negative control, keratinocytes were labeled with biotin-RQRR. Equal aliquots of cleared lysates<br />

were analyzed by Western blot for total matriptase, prior to streptavidin pull down of biotinylated<br />

proteins. Proteins were released from the streptavidin-coated resin by boiling in SDS sample<br />

buffer and analyzed by Western blotting. By labeling of keratinocytes isolated from mice<br />

expressing WT matriptase with biotin-RQRR-Cmk, we were able to detect active matriptase (Fig.<br />

6, lane 2). No matriptase was detected when labeling with a corresponding control peptide;<br />

biotin-RQRR (Fig. 6, lane 3). No matriptase could be detected in lysates or pull downs of<br />

keratinocytes from matriptase-deficient mice (Fig. 6, lanes 4-6 and 10-12), whereas matriptase<br />

was easily detected in all lysates of the 3 differently treated keratinocyte cultures from mice<br />

expressing WT matriptase (Fig. 6, lanes 7-9). Evaluation of total surface biotinylation and biotin-<br />

RQRR-Cmk labeling of WT murine keratinocytes showed that only a fraction of surface-associated<br />

matriptase on WT keratinocytes cells could be detected by means of biotin-RQRR-Cmk (compare<br />

Fig. 6 lanes 1 and 2). Thus, we have established an assay for detection of active matriptase that<br />

can be easily applied in other cell systems and other species.<br />

Discussion<br />

Matriptase is essential for postnatal survival and tight regulation of matriptase is crucial for a<br />

number of physiological functions including development and maintenance of epithelia [3-5;12-<br />

14;16;17]. Deregulated matriptase activity can also promote skin carcinogenesis [19]. It is<br />

therefore desirable to obtain methods to detect active matriptase to be able to assess active<br />

matriptase under these processes. This has proven difficult as no specific matriptase inhibitor<br />

and/or substrate is commercially available.<br />

For these reasons, we combined antibody specificity with the high affinity of biotin-streptavidin<br />

interaction in the design of a peptide inhibitor-based assay for detection of active matriptase. We


engineered a chloromethyl ketone based tetra-peptide inhibitor that allows for extraction of<br />

inhibitor bound matriptase from complex media by means of an N-terminal biotin moiety.<br />

Specificity of the assay is obtained by Western blot analysis employing specific antibodies against<br />

matriptase. Chloromethyl ketone based inhibitors have been extensively used as inhibitors of<br />

proteases and are active site-directed inhibitors that irreversible bind serine proteases by<br />

alkylation the active site histidine residue of serine proteases rendering it catalytic inactive [31].<br />

The assay is based on the assumption that active matriptase is equivalent to inhibitor bound<br />

matriptase. We show here that matriptase SPD activity is efficiently inhibited by biotin-RQRR-Cmk<br />

and that biotin-RQRR-Cmk is able to label active matriptase on the cell surface of cultured cells.<br />

Upon activation, matriptase rapidly forms a complex with HAI-1 [26;34;39]. HAI-1 is a Kunitz-type<br />

transmembrane serine protease inhibitor that binds to active matriptase in a reversible and<br />

competitive manner as is typical for this family of protease inhibitors [24;40]. The matriptase-HAI-<br />

1 complex comprises a docking of the Kunitz domain I of HAI-1 onto the active site and close<br />

surroundings thereby “capping” the substrate accessibility to the active site of the protease [35].<br />

We found that HAI-1 and biotin-RQRR-Cmk compete for active matriptase and although HAI-1<br />

binds matriptase in an irreversible, but yet very stable manner, biotin-RQRR-Cmk is not able to<br />

dissociate the matriptase-HAI-1 complex under the conditions used in the study. Moreover, the<br />

presence of biotin-RQRR-Cmk hinders matriptase-HAI-1 complex formation.<br />

In most epithelial cell lines and tissues a large fraction of matriptase is found in its zymogen form<br />

[23;26]. In accordance with this, we show that only a fraction of surface-associated matriptase<br />

could be labeled with biotin-RQRR-biotin from cell cultures of primary murine keratinocytes and<br />

unstimulated Caco-2 cells. The zymogen form of trypsin-like proteases is considered to exist in<br />

equilibrium between two conformations, an inactive conformation and an active conformation.<br />

Although most serine proteases are strongly in favor of the inactive form, intrinsic activity has<br />

been reported for a number of autoactivating proteases [36]. Matriptase is an autoactivating<br />

protease implying that the zymogen form has an inherent protease activity. We show that<br />

zymogen matriptase is able to bind biotin-RQRR-Cmk. This finding is supported by a recent study<br />

where a recombinant form of zymogen matriptase in an in vitro assay reacts with an inhibitor<br />

similar to the one used in the present study [41].<br />

Acid-induced activation of matriptase is ubiquitous among epithelial and carcinoma cells and is<br />

followed by rapid complex formation with HAI-1 [24;26;34]. Concurrent with this, we find that<br />

treatment of cells with physiological phosphate buffer pH 6.0 induces matriptase-HAI-1 complex<br />

formation in Caco-2 cells. We also find that a lowering of pH induces a rise in the level of<br />

matriptase that can be labeled with the inhibitor suggesting that there is a brief window between<br />

activation of matriptase and complex formation with HAI-1. This may be of physiological<br />

relevance, as the pericellular environment turns acidic in the transitional layer of the epidermis,<br />

the precise location where matriptase initiates a proteolytic cascade crucial for terminal<br />

epidermal differentiation [3;5;42].<br />

In summary, we have established an assay for detection active matriptase in cell culture. The<br />

availability of an assay for detection of active matriptase can greatly contribute to the further<br />

understanding of the complex biology of matriptase.


Figures<br />

Fig.1 Inhibition of matriptase SPD-mediated chromogenic substrate cleavage by biotin-RQRR-<br />

Cmk. (A) 0.2 μM matriptase SPD was incubated for 10 min at 37°C with (black triangle) or without<br />

(square) 50 μM biotin-RQRR-Cmk before addition the chromogenic substrate to a final<br />

concentration of 300μM. The reactivity of biotin-RQRR-Cmk was tested after 180 min of preincubation<br />

at 37°C (closed circle). Matriptase SPD was unable to cleave the chromogenic<br />

substrate, when biotin-RQRR-Cmk was added (black triangle) even after 180 min pre-incubation<br />

before addition to reaction buffer (closed circle). (B) 0.2nM matriptase SPD was incubated for 10<br />

min at 37°C with (black triangle) or without (square) 5 nM biotin-RQRR-Cmk before addition the<br />

chromogenic substrate to a final concentration of 300μM. The reactivity of biotin-RQRR-Cmk was<br />

tested with 60 min, 120 min, and 180 min of pre-incubation at 37°C before addition to reaction<br />

buffer. 5nM Biotin-RQRR-Cmk is able to completely inhibit substrate cleavage after 60 min preincubation<br />

at 37°C (open circle) and is still able to partly inhibit substrate cleavage after 180 min<br />

pre-incubation at 37°C (closed circle). All measurements were performed in 20 mM TRIS pH 7.4,<br />

140 mM NaCl supplemented with 0.1% BSA at 37°C. The reaction buffer was used as an internal<br />

control to verify that the peptidolytic activity originated from matriptase SPD (grey triangle) and<br />

not unspecific degradation of the substrate over time. Each plot shows the change in optical<br />

density at 405 nm of the reaction mixture as a function of reaction time. The presence of active<br />

protease results in a continued release of a yellow cleavage product resulting in a linear color<br />

development in agreement with a pseudo 1 st order reaction due to the high molar excess of<br />

substrate to protease. Results shown are representative of 3 independent experiments.


Fig. 2. Biotin-RRQR-Cmk reacts with a subset of matriptase molecules on the surface of Caco-2<br />

cells. 11 day post-confluent Caco-2 cells grown on Transwell filters were labeled with 50μM biotin-<br />

RQRR-Cmk from the basolateral side for the times indicated (2-180 min) at 37°C. As a measure of<br />

steady state levels of matriptase, cells were incubated on the basolateral membrane with S-NHS-<br />

SS-biotin at 4°C (lane 1) to biotinylate cell surface proteins. As a negative control, cells were<br />

labeled from the basolateral side with 50 μM control peptide; biotin-RQRR (lane 2). Biotinylated<br />

proteins were precipitated using streptavidin-coated resin and the streptavidin pull downs were<br />

analyzed by non-reducing SDS-PAGE and Western blotting using the monoclonal matriptase<br />

antibody; M32. A tenth of the surface biotinylated sample was loaded (lane 1); whereas total<br />

sample volume was loaded for the other samples (lanes 2-6). Surface biotinylation with S-NHS-SSbiotin<br />

resulted in a band of high intensity (lane 1), indicating that matriptase is abundantly present<br />

on the basolateral membrane. There is a clear time dependence of biotin-RQRR-Cmk binding to<br />

matriptase with weak detection of matriptase after 30 min to the most prominent band after 180<br />

min (lanes 3-6). No matriptase could be detected with the control peptide; biotin-RQRR (CTRL, lane<br />

2). Position of the molecular weight markers (kDa) is indicated on the left. Results shown are<br />

representative of 3 independent experiments.


Fig. 3: Matriptase activation, activity and inhibitor complex formation at pH 6.0. (A) 11 days<br />

post-confluence filter grown Caco-2 cells were either treated at pH 6.0 (lanes 2, 4, 6, and 8) or left<br />

untreated (Lanes 1, 3, 5, and 7) and lysates were analyzed by Western blotting with antibody<br />

against total matriptase (M24; lane 1 and 2), matriptase SPD (IM1014; lanes 5 and 6), matrpitase-<br />

HAI-1 complex (M69; lanes 3 and 4) and HAI-1 (lanes 7 and 8). Samples in lane 1-4, 7 and 8 were<br />

not boiled, while samples in lanes 5 and 6 were boiled and reduced to dissociate the S-S bridged<br />

SPD from the stem domain of activated matriptase. Non-boiled samples of untreated cells reveal<br />

that matriptase is primarily present as a prominent band of 70 kDa and as a minor band at 130<br />

kDa representing the matriptase-HAI-1 complex form (lane 1). Treatment at low pH reveals<br />

matriptase primarily in the 130 kDa complex form and a minor fraction in the 70 kDa form (lane<br />

2). The same rise in matriptase-HAI-1 complex can be detected with the M69 antibody (lanes 3 and<br />

4) and the HAI-1 antibody M19 (lanes 7 and 8). Under reducing conditions, matriptase can be<br />

detected mainly as the 70 kDa form, but also as two band of app. 30 kDa in size (lane 5). These 30<br />

kDa forms are more prominent than the 70 kDa form when cells were treated at pH 6.0 (lane 6).<br />

Treatment with phosphate buffer pH 6.0 and DTT is indicated by +/-. Position of the molecular<br />

weight markers (kDa) is indicated on the left. (B) 0.2 μM SPD were incubated for 10 min at 37°C<br />

with (black triangle) or without (square) 50 μM biotin-RQRR-Cmk before addition the chromogenic<br />

substrate to a final concentration of 300μM. All experiments were performed in 20 mM citric acid<br />

buffer pH 6.0, 140 mM NaCl and 0.1% BSA at 37°C. Matriptase SPD is able to cleave the<br />

chromogenic substrate resulting in color development of the sample at pH 6.0, when biotin-RQRR-<br />

Cmk was omitted (square) and an initial concentration of 50μM biotin-RQRR-Cmk is able to<br />

quench matriptase SPD activity towards the chromogenic substrate also at pH 6.0. A control<br />

containing only buffer verified that the peptidolytic activity originated from matriptase SPD (grey<br />

triangle) and not unspecific degradation of the substrate over time. Results shown are<br />

representative of 3 independent experiments.


Fig. 4. HAI-1 and biotin-RQRR-Cmk competes in inhibition of matriptase<br />

11 days post-confluent Caco-2 cells grown on Transwell filters were labeled with 50 μM biotin-<br />

RQRR-Cmk at neutral pH 7.4 (lane 3 and 7), in physiological phosphate buffer pH 6.0 (Lane 4 and<br />

8), or at neutral pH with a 30 min pre-incubation treatment with physiological phosphate buffer<br />

pH 6.0 (Lanes 5 and 9) for 30 min at 37°C. Samples of lysates were analyzed under non-boiled and<br />

non-reducing conditions (lanes 1-5). Labeled proteases were precipitated using streptavidin<br />

coated resin and released from the beads by boiling. As a negative control, lysate of cells treated<br />

with only physiological phosphate buffer pH 6.0 for 30 min was precipitated (CTRL, lanes 6). The<br />

streptavidin pull downs and non-boiled lysates of the same samples were analyzed by Western<br />

blotting using the monoclonal M32 antibody. Zymogen activation and matriptase-HAI-1 complex<br />

formation is induced with pH 6.0 treatment (lanes 1 and 2, respectively). Likewise, the amount of<br />

matriptase able to bind biotin-RQRR-Cmk also increases after pH 6.0 treatment (compare lanes 7<br />

and 8) and the presence of biotin-RQRR-Cmk hinders complex formation (compare lanes 4 and 5).<br />

Matriptase activation induced prior to biotin-RQRR-Cmk treatment result in matriptase-HAI-1<br />

complex formation (lane 5) and only low levels of matriptase binds biotin-RQRR-Cmk under this<br />

condition (lane 9) comparable to the level detected when treating cells with biotin-RQRR-Cmk at<br />

neutral pH (lane 7). Position of the molecular weight markers (kDa) is indicated on the left. Results<br />

shown are representative of 1 experiment.


Fig. 5. Biotin-RQRR-Cmk inhibits both matriptase zymogen and activation site cleaved<br />

matriptase. 11 days post-confluent Caco-2 cells grown on Transwell filters were labeled with 50<br />

μM biotin-RQRR-Cmk from the basolateral side for 180 min at 37°C. As a negative control, cells<br />

were labeled from the basolateral side with 50 μM control peptide; biotin-RQRR (CTRL), under the<br />

same conditions. Labeled proteases were precipitated using streptavidin-coated resin and the<br />

streptavidin pull downs were analyzed by reducing SDS-PAGE and Western blotting using the<br />

IM1014 antibody raised against matriptase SPD. The biotin-RQRR-biotin peptide labels both the 70<br />

kDa zymogen matriptase and activation site cleaved matriptase, which under these conditions can<br />

be detected as a 30 kDa band. Position of the molecular weight markers (kDa) is indicated on the<br />

left and position of matriptase zymogen and SPD is indicated on the right. Results shown are<br />

representative of 3 independent experiments.


Fig. 6: Detection of active matriptase in cultured primary murine keratinocytes. Murine<br />

keratinocytes were isolated from newborn wild-type (WT) or matriptase-deficient pups and<br />

plated on collagen-coated plastic. The cells were grown until sub-confluence and then labeled<br />

with the S-NHS-SS-biotin (B: lane 1, 4, 7, and 10), with 50 μM biotin-RQRR-Cmk (lane 2, 5, 8, and<br />

11), or with 50 μM control peptide; biotin-RQRR (lanes 3, 6, 9, and 12). Labeled proteases were<br />

precipitated using streptavidin-coated resin, and released from the beads by boiling and analyzed<br />

by SDS-PAGE and Western blotting using the matriptase antibody AF3946. Matriptase is detected<br />

in all lysates of WT keratinocytes (lanes 7-9), whereas no matriptase is detected in the lysates or<br />

pull downs from matriptase-deficient keratinocytes (lanes 4-6 and 10-12). The pull down fraction<br />

of S-NHS-SS-biotin labeled cells (lane 1) as well as the loaded aliquots of WT lysates (lanes 1-3 and<br />

7-9) shows that matriptase is abundantly present at the plasma membrane and that the fraction<br />

of peptidolytic active matriptase only constitute a small fraction of total surface-associated<br />

matriptase (lane 2). Results shown are representative of 2 independent experiments.


References<br />

[1] Oberst MD, Singh B, Ozdemirli M, Dickson RB, Johnson MD, & Lin CY (2003) Characterization of<br />

matriptase expression in normal human tissues. J. Histochem. Cytochem., 51, 1017-1025.<br />

[2] Oberst M, Anders J, Xie B, Singh B, Ossandon M, Johnson M, Dickson RB, & Lin CY (2001) Matriptase and<br />

HAI-1 are expressed by normal and malignant epithelial cells in vitro and in vivo. Am. J. Pathol., 158,<br />

1301-1311.<br />

[3] List K, Szabo R, Molinolo A, Nielsen BS, & Bugge TH (2006) Delineation of matriptase protein expression<br />

by enzymatic gene trapping suggests diverging roles in barrier function, hair formation, and squamous<br />

cell carcinogenesis. Am. J. Pathol., 168, 1513-1525.<br />

[4] List K, Kosa P, Szabo R, Bey AL, Wang CB, Molinolo A, & Bugge TH (2009) Epithelial integrity is<br />

maintained by a matriptase-dependent proteolytic pathway. Am. J. Pathol., 175, 1453-1463.<br />

[5] List K, Haudenschild CC, Szabo R, Chen W, Wahl SM, Swaim W, Engelholm LH, Behrendt N, & Bugge TH<br />

(2002) Matriptase/MT-SP1 is required for postnatal survival, epidermal barrier function, hair follicle<br />

development, and thymic homeostasis. Oncogene, 21, 3765-3779.<br />

[6] Basel-Vanagaite L, Attia R, Ishida-Yamamoto A, Rainshtein L, Ben AD, Lurie R, Pasmanik-Chor M,<br />

Indelman M, Zvulunov A, Saban S, Magal N, Sprecher E, & Shohat M (2007) Autosomal recessive<br />

ichthyosis with hypotrichosis caused by a mutation in ST14, encoding type II transmembrane serine<br />

protease matriptase. Am. J. Hum. Genet., 80, 467-477.<br />

[7] Avrahami L, Maas S, Pasmanik-Chor M, Rainshtein L, Magal N, Smitt J, van MJ, Shohat M, & Basel-<br />

Vanagaite L (2008) Autosomal recessive ichthyosis with hypotrichosis syndrome: further delineation of<br />

the phenotype. Clin. Genet., 74, 47-53.<br />

[8] Alef T, Torres S, Hausser I, Metze D, Tursen U, Lestringant GG, & Hennies HC (2009) Ichthyosis, follicular<br />

atrophoderma, and hypotrichosis caused by mutations in ST14 is associated with impaired profilaggrin<br />

processing. J. Invest Dermatol., 129, 862-869.<br />

[9] Sales KU, Masedunskas A, Bey AL, Rasmussen AL, Weigert R, List K, Szabo R, Overbeek PA, & Bugge TH<br />

(2010) Matriptase initiates activation of epidermal pro-kallikrein and disease onset in a mouse model of<br />

Netherton syndrome. Nat. Genet., 42, 676-683.<br />

[10] List K, Szabo R, Wertz PW, Segre J, Haudenschild CC, Kim SY, & Bugge TH (2003) Loss of proteolytically<br />

processed filaggrin caused by epidermal deletion of Matriptase/MT-SP1. J. Cell Biol., 163, 901-910.<br />

[11] Tanaka H, Nagaike K, Takeda N, Itoh H, Kohama K, Fukushima T, Miyata S, Uchiyama S, Uchinokura S,<br />

Shimomura T, Miyazawa K, Kitamura N, Yamada G, & Kataoka H (2005) Hepatocyte growth factor<br />

activator inhibitor type 1 (HAI-1) is required for branching morphogenesis in the chorioallantoic<br />

placenta. Mol. Cell Biol., 25, 5687-5698.<br />

[12] Szabo R, Hobson JP, Christoph K, Kosa P, List K, & Bugge TH (2009) Regulation of cell surface protease<br />

matriptase by HAI2 is essential for placental development, neural tube closure and embryonic survival in<br />

mice. Development, 136, 2653-2663.<br />

[13] Szabo R, Molinolo A, List K, & Bugge TH (2007) Matriptase inhibition by hepatocyte growth factor<br />

activator inhibitor-1 is essential for placental development. Oncogene, 26, 1546-1556.


[14] Carney TJ, von der HS, Sonntag C, Amsterdam A, Topczewski J, Hopkins N, & Hammerschmidt M (2007)<br />

Inactivation of serine protease Matriptase1a by its inhibitor Hai1 is required for epithelial integrity of<br />

the zebrafish epidermis. Development, 134, 3461-3471.<br />

[15] Nagaike K, Kawaguchi M, Takeda N, Fukushima T, Sawaguchi A, Kohama K, Setoyama M, & Kataoka H<br />

(2008) Defect of hepatocyte growth factor activator inhibitor type 1/serine protease inhibitor, Kunitz<br />

type 1 (Hai-1/Spint1) leads to ichthyosis-like condition and abnormal hair development in mice. Am. J.<br />

Pathol., 173, 1464-1475.<br />

[16] Szabo R, Kosa P, List K, & Bugge TH (2009) Loss of matriptase suppression underlies spint1 mutationassociated<br />

ichthyosis and postnatal lethality. Am. J. Pathol., 174, 2015-2022.<br />

[17] Mathias JR, Dodd ME, Walters KB, Rhodes J, Kanki JP, Look AT, & Huttenlocher A (2007) Live imaging of<br />

chronic inflammation caused by mutation of zebrafish Hai1. J. Cell Sci., 120, 3372-3383.<br />

[18] Szabo R, Uzzun SK, Kosa P, Shylo NA, Godiksen S, Hansen KK, Friis S, Gutkind JS, Vogel LK, Hummler E,<br />

Camerer E, & Bugge TH (2012) Reduced Prostasin (CAP1/PRSS8) Activity Eliminates HAI-1 and HAI-2<br />

Deficiency-Associated Developmental Defects by Preventing Matriptase Activation. PLoS. Genet., 8,<br />

e1002937.<br />

[19] List K, Szabo R, Molinolo A, Sriuranpong V, Redeye V, Murdock T, Burke B, Nielsen BS, Gutkind JS, &<br />

Bugge TH (2005) Deregulated matriptase causes ras-independent multistage carcinogenesis and<br />

promotes ras-mediated malignant transformation. Genes Dev., 19, 1934-1950.<br />

[20] Szabo R, Rasmussen AL, Moyer AB, Kosa P, Schafer JM, Molinolo AA, Gutkind JS, & Bugge TH (2011) c-<br />

Met-induced epithelial carcinogenesis is initiated by the serine protease matriptase. Oncogene.<br />

[21] Kosa P, Szabo R, Molinolo AA, & Bugge TH (2011) Suppression of Tumorigenicity-14, encoding<br />

matriptase, is a critical suppressor of colitis and colitis-associated colon carcinogenesis. Oncogene.<br />

[22] Chen YW, Wang JK, Chou FP, Chen CY, Rorke EA, Chen LM, Chai KX, Eckert RL, Johnson MD, & Lin CY<br />

(2010) Regulation of the matriptase-prostasin cell surface proteolytic cascade by hepatocyte growth<br />

factor activator inhibitor-1 during epidermal differentiation. J. Biol. Chem., 285, 31755-31762.<br />

[23] Lee MS, Kiyomiya K, Benaud C, Dickson RB, & Lin CY (2005) Simultaneous activation and hepatocyte<br />

growth factor activator inhibitor 1-mediated inhibition of matriptase induced at activation foci in human<br />

mammary epithelial cells. Am. J. Physiol Cell Physiol, 288, C932-C941.<br />

[24] Benaud C, Dickson RB, & Lin CY (2001) Regulation of the activity of matriptase on epithelial cell surfaces<br />

by a blood-derived factor. Eur. J. Biochem., 268, 1439-1447.<br />

[25] Friis S, Godiksen S, Bornholdt J, Selzer-Plon J, Rasmussen HB, Bugge TH, Lin CY, & Vogel LK (2010)<br />

Transport via the transcytotic pathway makes prostasin available as a substrate for matriptase. J. Biol.<br />

Chem..<br />

[26] Lee MS, Tseng IC, Wang Y, Kiyomiya K, Johnson MD, Dickson RB, & Lin CY (2007) Autoactivation of<br />

matriptase in vitro: requirement for biomembrane and LDL receptor domain. Am. J. Physiol Cell Physiol,<br />

293, C95-105.<br />

[27] Yuan C, Chen L, Meehan EJ, Daly N, Craik DJ, Huang M, & Ngo JC (2011) Structure of catalytic domain of<br />

Matriptase in complex with Sunflower trypsin inhibitor-1. BMC. Struct. Biol., 11, 30.


[28] Lichti U, Anders J, & Yuspa SH (2008) Isolation and short-term culture of primary keratinocytes, hair<br />

follicle populations and dermal cells from newborn mice and keratinocytes from adult mice for in vitro<br />

analysis and for grafting to immunodeficient mice. Nat. Protoc., 3, 799-810.<br />

[29] Lin CY, Anders J, Johnson M, & Dickson RB (1999) Purification and characterization of a complex<br />

containing matriptase and a Kunitz-type serine protease inhibitor from human milk. J. Biol. Chem., 274,<br />

18237-18242.<br />

[30] Beliveau F, Desilets A, & Leduc R (2009) Probing the substrate specificities of matriptase, matriptase-2,<br />

hepsin and DESC1 with internally quenched fluorescent peptides. FEBS J., 276, 2213-2226.<br />

[31] Powers JC, Asgian JL, Ekici OD, & James KE (2002) Irreversible inhibitors of serine, cysteine, and<br />

threonine proteases. Chem. Rev., 102, 4639-4750.<br />

[32] Anderson PJ & Bock PE (2001) Biotin derivatives of D-Phe-Pro-Arg-CH2Cl for active-site-specific labeling<br />

of thrombin and other serine proteinases. Anal. Biochem., 296, 254-261.<br />

[33] Buzza MS, Netzel-Arnett S, Shea-Donohue T, Zhao A, Lin CY, List K, Szabo R, Fasano A, Bugge TH, &<br />

Antalis TM (2010) Membrane-anchored serine protease matriptase regulates epithelial barrier<br />

formation and permeability in the intestine. Proc. Natl. Acad. Sci. U. S. A.<br />

[34] Tseng IC, Xu H, Chou FP, Li G, Vazzano AP, Kao JP, Johnson MD, & Lin CY (2010) Matriptase activation, an<br />

early cellular response to acidosis. J. Biol. Chem., 285, 3261-3270.<br />

[35] Friedrich R, Fuentes-Prior P, Ong E, Coombs G, Hunter M, Oehler R, Pierson D, Gonzalez R, Huber R,<br />

Bode W, & Madison EL (2002) Catalytic domain structures of MT-SP1/matriptase, a matrix-degrading<br />

transmembrane serine proteinase. J. Biol. Chem., 277, 2160-2168.<br />

[36] Williams EB, Krishnaswamy S, & Mann KG (1989) Zymogen/enzyme discrimination using peptide<br />

chloromethyl ketones. J. Biol. Chem., 264, 7536-7545.<br />

[37] Ranby M, Bergsdorf N, & Nilsson T (1982) Enzymatic properties of the one- and two-chain form of tissue<br />

plasminogen activator. Thromb. Res., 27, 175-183.<br />

[38] Oberst MD, Williams CA, Dickson RB, Johnson MD, & Lin CY (2003) The activation of matriptase requires<br />

its noncatalytic domains, serine protease domain, and its cognate inhibitor. J. Biol. Chem., 278, 26773-<br />

26779.<br />

[39] Benaud C, Oberst M, Hobson JP, Spiegel S, Dickson RB, & Lin CY (2002) Sphingosine 1-phosphate,<br />

present in serum-derived lipoproteins, activates matriptase. J. Biol. Chem., 277, 10539-10546.<br />

[40] Laskowski M, Jr. & Kato I (1980) Protein inhibitors of proteinases. Annu. Rev. Biochem., 49, 593-626.<br />

[41] Inouye K, Yasumoto M, Tsuzuki S, Mochida S, & Fushiki T (2010) The optimal activity of a<br />

pseudozymogen form of recombinant matriptase under the mildly acidic pH and low ionic strength<br />

conditions. J. Biochem., 147, 485-492.<br />

[42] Netzel-Arnett S, Currie BM, Szabo R, Lin CY, Chen LM, Chai KX, Antalis TM, Bugge TH, & List K (2006)<br />

Evidence for a matriptase-prostasin proteolytic cascade regulating terminal epidermal differentiation. J.<br />

Biol. Chem., 281, 32941-32945.


Discussion and perspectives<br />

In the following section the data presented in the three manuscripts enclosed in this thesis will be<br />

discussed starting with a short summary of the key findings<br />

In paper I, we delineate the subcellular trafficking of matriptase, prostasin and HAI-1 to address<br />

how matriptase can act as an upstream activator of prostasin in the epidermis, despite their<br />

different subcellular localizations in polarized epithelial cells. We show that matriptase is routed<br />

to the basolateral plasma membrane, where it is cleaved into its active form. We also found that<br />

both prostasin and HAI-1 are transported to the basolateral plasma membrane from where they<br />

are internalized and transcytosed to the apical membrane. No transcytosis of matriptase could be<br />

detected although matriptase is endocytosed in a complex with HAI-1 from the basolateral<br />

plasma membrane. Interestingly, we could detect active prostasin on both the apical and the<br />

basolateral plasma membrane, as well as active matriptase on the basolateral plasma membrane<br />

by their ability to bind inhibitor-coupled beads. This study propose the basolateral plasma<br />

membrane as the site for matriptase-prostasin interaction and hereby suggest how the apically<br />

located substrate, prostasin, can be cleaved and activated by a basolateral localized protease,<br />

matriptase, in terminal epidermal differentiation.<br />

To gain more knowledge of matriptase-mediated proteolysis in normal physiological processes,<br />

we performed genetic epistasis analyses to identify new component(s) of matriptase-dependent<br />

proteolytic pathways critical to embryogenesis (paper II). We found that a hypomorphic mutation<br />

in Prss8, the gene encoding prostasin, in a similar manner to ablation of ST14, encoding<br />

matriptase, restored nearly all developmental defects associated with HAI-1 and HAI-2 deficiency<br />

indicating that matriptase and prostasin function in the same pathway(s) crucial for survival of<br />

newborn mice. Interestingly, we show that activation of matriptase does not occur in prostasin<br />

deficient placental tissues, whereas activated prostasin is readily detectable in matriptase<br />

deficient placental tissues. Thus, contrary to studies of the epidermis, we demonstrate that<br />

prostasin acts upstream of matriptase and is requires for its activation in placental tissues.<br />

In an attempt to obtain tools for detection of active matriptase, we designed a chloromethyl<br />

ketone peptide inhibitor of matriptase and were hereby able to optimize conditions and establish<br />

an assay for detection of active matriptase in live cell cultures. We found that matriptase is active<br />

on the basolateral plasma membrane by its ability to bind the chloromethyl ketone peptide<br />

inhibitor, biotin-RQRR-Cmk, however only a fraction of total surface matriptase could by extracted<br />

by means of the inhibitor. We show that the chloromethyl ketone peptide inhibitor competes<br />

with HAI-1 for binding to matriptase and that the matriptase zymogen has an intrinsic activity that<br />

enables it to bind the chloromethyl ketone peptide inhibitor. Moreover, we are able to modify the<br />

assay to confirm the presence of active matriptase in cultures of primary murine keratinocytes.<br />

Within recent years, the conception of a matriptase-prostasin proteolytic pathway has been<br />

established in epidermal homeostasis. The existence of such a cascade is based on the facts that<br />

active prostasin is absent in matriptase-deficient epidermis, and that matriptase-deficient mice<br />

85


and mice deficient of prostasin in the skin display nearly indistinguishable phenotypes. Also the<br />

two proteases are co-expressed in a number of epithelial tissues besides the epidermis<br />

[4;9;12;15;16;116].<br />

We have previously delineated the intracellular transport of human HAI-1 recombinantly<br />

expressed in a different model of the polarized epithelia, namely the MDCK cell line. This study<br />

show that HAI-1 has a complex subcellular itinerary and that HAI-1 is transcytosed from the<br />

basolateral plasma membrane to the apical membrane (supplementary I; [27]). The fact that<br />

matriptase in complex with HAI-1 has been purified from human milk, and unpublished results<br />

showing that a 80 kDa protein was co-precipitated with HAI-1 lead us to the hypothesis that HAI-<br />

1, in addition to its protease inhibitor function, could play a role in transporting matriptase in a<br />

matriptase–HAI-1 complex from the basolateral plasma membrane to the apical plasma<br />

membrane, where matriptase could activate prostasin (Sine Godiksen, unpublished results;<br />

supplementary I [27], [44]).<br />

In the current studies (paper I and III), we have used the human Caco-2 cell line as a model for the<br />

polarized epithelial cell. Paper I reports a polarized distribution of endogenous expressed<br />

matriptase, prostasin and HAI-1 in Caco-2 cells, which is in accordance to findings in other cell<br />

types and tissues, showing matriptase and HAI-1 on the basolateral plasma membrane and<br />

prostasin on the apical plasma membrane [9;11;21;23-26]. This makes the Caco-2 cell line a<br />

suitable model system for the studies conducted in paper I and paper III.<br />

The results obtained in paper I also confirmed the previous results obtained in MDCK cells on<br />

subcellular trafficking of HAI-1. However, we were unable to detect transcytosis of matriptase<br />

from the basolateral to the apical plasma membrane as hypothesized, although shedding of HAI-<br />

1-complexed matriptase to the apical media has been observed in polarized Caco-2 cells by us and<br />

others (Stine Friis, unpublished result; [78]).<br />

Instead the results from paper I point to a mechanism where prostasin is routed to the basolateral<br />

plasma membrane followed by transcytosis to the apical membrane as an activated protease.<br />

Thus, the complex itinerary of prostasin, position matriptase and prostasin on the same<br />

subcellular domain and explain how matriptase can activate prostasin in terminal epidermal<br />

differentiation. Whether matriptase act as a direct activator of prostasin in Caco-2 cells, or if other<br />

proteases are responsible for the activation is not shown. On the other hand, our proposed<br />

mechanism for the matriptase-prostasin interaction does not specify any order of the interaction<br />

between the two proteases.<br />

With the generation of matriptase-deficient mice it became clear that matriptase plays a role in<br />

global epithelial development and maintenance [4;5]. Matriptase is believed to be at the pinnacle<br />

of protease cascade(s) because of its ability to autoactivate [28;29;91].<br />

However, it is still unclear which pathways and components that are involved in matriptasedependent<br />

proteolysis. Matriptase and prostasin are co-expressed in a variety of epithelia besides<br />

the epidermis, which has led to speculations of a role for the matriptase-prostasin proteolytic<br />

cascade in other epithelia [9].<br />

In genetic analyses to identify new components of matriptase-dependent proteolytic pathway(s)<br />

necessary for survival of newborn mice, we identified prostasin to be critical to all matriptaseinduced<br />

embryonic defects in both HAI-1 and HAI-2 deficient mice (paper II). Paradoxically, our<br />

study reports that in placental tissue prostasin acts upstream of matriptase and is required for the<br />

86


conversion of zymogen matriptase to active matriptase, which is contrary to what is previously<br />

described for epidermis [15-17;21]. To consolidate this in vivo finding, we show that prostasin is<br />

able to efficiently cleave and activate zymogen matriptase in a recombinant cell based assay.<br />

Interestingly, our data is supported by cell culture studies where addition of active prostasin is<br />

sufficient to activate zymogen matriptase expressed recombinantly in KOLF cells as well as in<br />

HaCaT cells that have an endogenous expression of matriptase [57].<br />

Fig. 13. Outline of matriptase-prostasin interaction in the epidermis and prostasin-matriptase<br />

interaction in placental tissue.<br />

In the epidermis, matriptase acts upstream of prostasin and is required for its action. In placental<br />

tissue the relationship between the two proteases is reversed; thus prostasin is the upstream<br />

protease and required for matriptase activation.<br />

Thus, prostasin may act both as a downstream target but also as an upstream activator of<br />

matriptase in a tissue specific manner, see fig. 13, and/or perhaps constitute a positive feedback<br />

loop similar to the amplification protease cascades operating in coagulation [153]. Further studies<br />

in different epithelial tissues are highly desirable to determine the tissue-specific interplay<br />

between the two proteases, and to identify new components and pathways dependent on<br />

matriptase proteolytic activity in global epithelial homeostasis. It would be interesting to assess<br />

the presence of active matriptase and active prostasin in prostasin-deficient and matriptase-<br />

87


deficient intestinal tissue, respectively, as this is the epithelial tissue to which Caco-2 cells has the<br />

highest degree of resemblance. Still regardless of whether matriptase activates prostasin or vice<br />

versa, the fact that both active matriptase and active prostasin locate on the basolateral plasma<br />

membrane of Caco-2 cells as outlined in paper I can still apply to explain how two proteases with<br />

different steady state localization may interact. Thus it is feasible that in placental tissue,<br />

prostasin could, upon activation on the basolateral plasma membrane, activate matriptase before<br />

being transcytosed to the apical plasma membrane.<br />

The finding that prostasin act upstream of matriptase and is required for its activation is puzzling<br />

in regards to the well-documented ability of matriptase to autoactivate and the fact that prostasin<br />

is incapable of autoactivation due to an unfavorable isoleucine in the P1' position of the activation<br />

cleavage motif of prostasin [107]. It also raises the obvious question of the upstream activator of<br />

prostasin in this setting. Another activator of prostasin is the membrane-anchored serine<br />

protease hepsin [65]. However, hepsin deficient mice exhibit normal embryogenesis and adult<br />

mice display no aberrant phenotype suggesting that another protease is involved in prostasin<br />

activation [154;155]. Establishment of biological functions of the rest of the members of the<br />

family of membrane-anchored serine proteases could provide important information for<br />

determining an upstream activator of prostasin during embryogenesis.<br />

Another interesting aspect and additional layer of complexity of matriptase-prostasin interplay<br />

comes from studies on ENaC activation. Both matriptase and prostasin are able to activate ENaC<br />

in model systems [103;120]. Although membrane anchorage of prostasin is needed for activation<br />

of ENaC in xenopus oocytes, the catalytic activity of prostasin has been reported to be<br />

dispensable for the activational cleavage of ENaC [123;124;156]. This could suggest a role for<br />

prostasin as a co-factor important for the proteolytic activity of another protease towards ENaC.<br />

One could speculate that matriptase participate in prostasin-dependent activation of ENaC. In<br />

paper I we were able to detect matriptase on the apical membrane, and we and others have<br />

observed shedding of matriptase into the apical media, as well as matriptase is present in human<br />

milk (unpublished results, Stine Friis; [44;78]). These findings all indicate that matriptase in fact is<br />

present on the apical plasma membrane. Moreover, by means of the assay established in paper<br />

III, we have shown that active matriptase is present on the apical membrane (unpublished results,<br />

Sine Godiksen). This could position matriptase on the apical plasma membrane to activate ENaC in<br />

a prostasin-dependent manner.<br />

Perhaps prostasin could have a general role as a co-factor, as prostasin has been shown to<br />

enhances matriptase cleavage of EGFR in FT293 cells with no direct cleavage of EGFR by active<br />

prostasin [65]. In this regard, it would also be interesting to examine the necessity for prostasin<br />

catalytic activity in the dependence of GPI-anchored prostasin in plasmin-stimulated ENaC<br />

activation [157]. The role of prostasin as a co-factor could be addressed by the generation of<br />

tissue-specific knock in mouse models of catalytic dead prostasin by mutation of the active site<br />

serine. Thus, the association between matriptase and prostasin is complex and a great amount of<br />

work is needed to unravel the interrelationship of the two proteases in different tissues and cells<br />

types.<br />

88


This thesis also aimed at detecting active matriptase and determining in what subcellular<br />

compartment matriptase is proteolytically processed and activated, and thus able to cleave<br />

downstream substrate(s). In most cell lysates and tissue extracts, matriptase is found in either its<br />

zymogen form or in a complex with its inhibitor HAI‐1 [30;98]. In cell cultures of epithelial and<br />

carcinoma origin, matriptase activation can be induced under slight acidic conditions. This acidinduced<br />

activation of matriptase is fast and efficient and is followed by HAI-1 inhibition within few<br />

minutes [30;31;98;99]. This intimate link between activation and inhibition of matriptase leaves a<br />

small window of action for the protease to cleave its substrates and poses a challenge in detecting<br />

matriptase in its free active form.<br />

Different methods have been used in the three enclosed manuscripts to assess activation of<br />

matriptase and active matriptase. In paper II, we assay for the presence of active matriptase and<br />

active prostasin by co-IP with HAI-1. However, this does not give an assessment of active<br />

protease, but merely inform of whether activation site cleavage has occurred, as we precipitate<br />

HAI-1-complexed prostasin and HAI-1-complexed matriptase. In paper I, we use inhibitor-coupled<br />

sepharose to assess the level of active matriptase. The non-inhibitory LDLR domain of HAI-1 is<br />

important for the matriptase-HAI-1 interaction [28]. For this reason, we used non-endogenous<br />

inhibitors of matriptase and prostasin to avoid possible non-inhibitory interactions and thus to<br />

target only free active matriptase and prostasin. A shortcoming of this procedure is that the<br />

interaction between matriptase and the inhibitor-coupled beads occurs in cell lysates, as<br />

matriptase is sensitive to autoactivation in cell free systems [29].<br />

This prompted us to establish an assay for the detection of active matriptase in live cell cultures<br />

(paper III). The difficulties in setting up a specific assay for the detection of matriptase come from<br />

the great substrate overlap within this family and the absence of a specific inhibitor and/or<br />

specific substrate of matriptase [48;50]. We have based our assay on a chloromethyl ketone<br />

peptide inhibitor with a tetra peptide sequence based on a preferred substrate sequence of<br />

matriptase [50]. CMK peptide inhibitors have long been used in protease research and one of the<br />

great advantages is that they are active site inhibitors and bind proteases in a covalent manner by<br />

alkylation of the active site histidine [158;159]. Although some specificity can be obtained by<br />

altering the peptide sequence of this type of inhibitor, it is very difficult to get absolute specificity<br />

towards a single protease [158]. To circumvent the issue of substrate overlap, we employ specific<br />

antibodies and Western blot analysis in the detection of active matriptase.<br />

We show that active matriptase is present on the basolateral plasma membrane but only<br />

constitute a minor fraction of total surface-associated matriptase although we use vast amounts<br />

of chloromethyl ketone peptide inhibitor (paper III). However, the absence of a linker arm in our<br />

design of chloromethyl ketone peptide inhibitor could cause a reduced affinity for streptavidin<br />

and hereby a reduced detection of biotin-RQRR-Cmk labeled matriptase. For the serine protease<br />

thrombin, a similar assay set-up showed that a linker arm of 7-14 atoms produced the most<br />

sensitive detection in Western blotting [160].<br />

Moreover, our results indicate that biotin-RQRR-Cmk competes with HAI-1 for matriptase binding<br />

however; the precipitation step used prevents us from directly assessing this. Instead, applying<br />

monomeric avidin that bind biotin in a reversible manner would allow us to examine this.<br />

Interestingly, we also find that the zymogen matriptase has intrinsic activity. This is not an<br />

uncommon feature of serine proteases and has been shown for i.a. single-chain tPA [161]. The<br />

89


increase in catalytic activity after zymogen activation varies widely among the different members<br />

of the trypsin protease family, however for single-chain and two-chain tPA the catalytic activities<br />

vary by a factor of only 3-9 [161-165]. It would be interestingly to determine the catalytic capacity<br />

of zymogen matriptase; if strong the zymogen form of matriptase could perhaps be the biological<br />

relevant form in light of the intimate link between activation site cleavage of matriptase and<br />

subsequent inhibition by HAI-1.<br />

Whereas expression studies give valuable insights on matriptase localization and matriptase<br />

mRNA/protein levels, these findings does not reflect the biological active state of the protease.<br />

Having an assay that enables us to assess the level of active matriptase would be valuable in<br />

determining the activation degree of matriptase under different physiological and pathological<br />

states. This is particular the case in assessment of matriptase´s contribution in progression of<br />

cancer. However, applying the assay presented in paper III on tissue extracts would require some<br />

adjustments and also present challenges as matriptase is prone to spontaneous autoactivation in<br />

cell free systems [29]. Additionally it would be desirable with a specific molecular probe for active<br />

matriptase that would also allow visualization of matriptase activity in pathological conditions.<br />

There have been other attempts to identify inhibitors of matriptase. These inhibitors include bisbenzamidines,<br />

sun flower seed trypsin inhibitor-1 derived peptides, antibody-derived inhibitors,<br />

and small molecule peptidyl- derivatives inhibitors [149;166-175]. However, not all inhibitors are<br />

specific for matriptase, e.g. the chemical inhibitor (CVS-3983) of matriptase that was reported to<br />

suppress the prostate cancer growth in nude mice has a high selectivity but is not specific for<br />

matriptase [167]. On the other hand, antibodies constructs have been used for detection of<br />

matriptase positive cancer cells in a mouse xenografts model [166]. Recently an assay was<br />

reported that measured the activity of matriptase on epithelial airway cells. This study use a<br />

different approach than ours and assign proteolytic activity to matriptase by the difference in<br />

measured activity with and without addition of a non-commercial matriptase inactivating<br />

antibody [149]. Even so, despite the methods described above there is still a shortage for specific<br />

biochemical assays for detection of active matriptase.<br />

This thesis reports how complex subcellular trafficking of prostasin enables this protease to<br />

interact with matriptase on the basolateral plasma membrane of polarized epithelial cells and<br />

consolidate how HAI-1 can function as an inhibitor of both proteases. We have shown the<br />

subcellular location for matriptase activation and established an assay for detection of active<br />

matriptase on the surface of living cells. Moreover, this thesis describes a prostasin-matriptase<br />

cascade important for morphogenesis in mice. Together the data presented here has improved<br />

our understanding of matriptase´s role in epithelial physiology.<br />

90


References<br />

[1] Vine AK (2009) Recent advances in haemostasis and thrombosis. Retina, 29, 1-7.<br />

[2] Szabo R, Wu Q, Dickson RB, Netzel-Arnett S, Antalis TM, & Bugge TH (2003) Type II<br />

transmembrane serine proteases. Thromb. Haemost., 90, 185-193.<br />

[3] Szabo R & Bugge TH (2011) Membrane-anchored serine proteases in vertebrate cell and<br />

developmental biology. Annu. Rev. Cell Dev. Biol., 27, 213-235.<br />

[4] List K, Haudenschild CC, Szabo R, Chen W, Wahl SM, Swaim W, Engelholm LH, Behrendt N, &<br />

Bugge TH (2002) Matriptase/MT-SP1 is required for postnatal survival, epidermal barrier<br />

function, hair follicle development, and thymic homeostasis. Oncogene, 21, 3765-3779.<br />

[5] List K, Kosa P, Szabo R, Bey AL, Wang CB, Molinolo A, & Bugge TH (2009) Epithelial integrity is<br />

maintained by a matriptase-dependent proteolytic pathway. Am. J. Pathol., 175, 1453-1463.<br />

[6] List K, Szabo R, Molinolo A, Sriuranpong V, Redeye V, Murdock T, Burke B, Nielsen BS,<br />

Gutkind JS, & Bugge TH (2005) Deregulated matriptase causes ras-independent multistage<br />

carcinogenesis and promotes ras-mediated malignant transformation. Genes Dev., 19, 1934-<br />

1950.<br />

[7] Szabo R, Rasmussen AL, Moyer AB, Kosa P, Schafer JM, Molinolo AA, Gutkind JS, & Bugge TH<br />

(2011) c-Met-induced epithelial carcinogenesis is initiated by the serine protease matriptase.<br />

Oncogene.<br />

[8] Kosa P, Szabo R, Molinolo AA, & Bugge TH (2011) Suppression of Tumorigenicity-14, encoding<br />

matriptase, is a critical suppressor of colitis and colitis-associated colon carcinogenesis.<br />

Oncogene.<br />

[9] List K, Hobson JP, Molinolo A, & Bugge TH (2007) Co-localization of the channel activating<br />

protease prostasin/(CAP1/PRSS8) with its candidate activator, matriptase. J. Cell Physiol, 213,<br />

237-245.<br />

[10] Oberst M, Anders J, Xie B, Singh B, Ossandon M, Johnson M, Dickson RB, & Lin CY (2001)<br />

Matriptase and HAI-1 are expressed by normal and malignant epithelial cells in vitro and in<br />

vivo. Am. J. Pathol., 158, 1301-1311.<br />

[11] Oberst MD, Singh B, Ozdemirli M, Dickson RB, Johnson MD, & Lin CY (2003) Characterization<br />

of matriptase expression in normal human tissues. J. Histochem. Cytochem., 51, 1017-1025.<br />

[12] List K, Szabo R, Molinolo A, Nielsen BS, & Bugge TH (2006) Delineation of matriptase protein<br />

expression by enzymatic gene trapping suggests diverging roles in barrier function, hair<br />

formation, and squamous cell carcinogenesis. Am. J. Pathol., 168, 1513-1525.<br />

[13] Basel-Vanagaite L, Attia R, Ishida-Yamamoto A, Rainshtein L, Ben AD, Lurie R, Pasmanik-Chor<br />

M, Indelman M, Zvulunov A, Saban S, Magal N, Sprecher E, & Shohat M (2007) Autosomal<br />

recessive ichthyosis with hypotrichosis caused by a mutation in ST14, encoding type II<br />

transmembrane serine protease matriptase. Am. J. Hum. Genet., 80, 467-477.<br />

91


[14] Avrahami L, Maas S, Pasmanik-Chor M, Rainshtein L, Magal N, Smitt J, van MJ, Shohat M, &<br />

Basel-Vanagaite L (2008) Autosomal recessive ichthyosis with hypotrichosis syndrome:<br />

further delineation of the phenotype. Clin. Genet., 74, 47-53.<br />

[15] Alef T, Torres S, Hausser I, Metze D, Tursen U, Lestringant GG, & Hennies HC (2009)<br />

Ichthyosis, follicular atrophoderma, and hypotrichosis caused by mutations in ST14 is<br />

associated with impaired profilaggrin processing. J. Invest Dermatol., 129, 862-869.<br />

[16] Netzel-Arnett S, Currie BM, Szabo R, Lin CY, Chen LM, Chai KX, Antalis TM, Bugge TH, & List K<br />

(2006) Evidence for a matriptase-prostasin proteolytic cascade regulating terminal epidermal<br />

differentiation. J. Biol. Chem., 281, 32941-32945.<br />

[17] List K, Currie B, Scharschmidt TC, Szabo R, Shireman J, Molinolo A, Cravatt BF, Segre J, &<br />

Bugge TH (2007) Autosomal ichthyosis with hypotrichosis syndrome displays low matriptase<br />

proteolytic activity and is phenocopied in ST14 hypomorphic mice. J. Biol. Chem., 282, 36714-<br />

36723.<br />

[18] Desilets A, Beliveau F, Vandal G, McDuff FO, Lavigne P, & Leduc R (2008) Mutation G827R in<br />

matriptase causing autosomal recessive ichthyosis with hypotrichosis yields an inactive<br />

protease. J. Biol. Chem., 283, 10535-10542.<br />

[19] List K, Szabo R, Wertz PW, Segre J, Haudenschild CC, Kim SY, & Bugge TH (2003) Loss of<br />

proteolytically processed filaggrin caused by epidermal deletion of Matriptase/MT-SP1. J. Cell<br />

Biol., 163, 901-910.<br />

[20] Szabo R, Kosa P, List K, & Bugge TH (2009) Loss of matriptase suppression underlies spint1<br />

mutation-associated ichthyosis and postnatal lethality. Am. J. Pathol., 174, 2015-2022.<br />

[21] Chen YW, Wang JK, Chou FP, Chen CY, Rorke EA, Chen LM, Chai KX, Eckert RL, Johnson MD, &<br />

Lin CY (2010) Regulation of the matriptase-prostasin cell surface proteolytic cascade by<br />

hepatocyte growth factor activator inhibitor-1 during epidermal differentiation. J. Biol.<br />

Chem., 285, 31755-31762.<br />

[22] Carney TJ, von der HS, Sonntag C, Amsterdam A, Topczewski J, Hopkins N, & Hammerschmidt<br />

M (2007) Inactivation of serine protease Matriptase1a by its inhibitor Hai1 is required for<br />

epithelial integrity of the zebrafish epidermis. Development, 134, 3461-3471.<br />

[23] Verghese GM, Gutknecht MF, & Caughey GH (2006) Prostasin regulates epithelial monolayer<br />

function: cell-specific Gpld1-mediated secretion and functional role for GPI anchor. Am. J.<br />

Physiol Cell Physiol, 291, C1258-C1270.<br />

[24] Kataoka H, Suganuma T, Shimomura T, Itoh H, Kitamura N, Nabeshima K, & Koono M (1999)<br />

Distribution of hepatocyte growth factor activator inhibitor type 1 (HAI-1) in human tissues.<br />

Cellular surface localization of HAI-1 in simple columnar epithelium and its modulated<br />

expression in injured and regenerative tissues. J. Histochem. Cytochem., 47, 673-682.<br />

[25] Kataoka H, Meng JY, Itoh H, Hamasuna R, Shimomura T, Suganuma T, & Koono M (2000)<br />

Localization of hepatocyte growth factor activator inhibitor type 1 in Langhans' cells of<br />

human placenta. Histochem. Cell Biol., 114, 469-475.<br />

92


[26] Satomi S, Yamasaki Y, Tsuzuki S, Hitomi Y, Iwanaga T, & Fushiki T (2001) A role for<br />

membrane-type serine protease (MT-SP1) in intestinal epithelial turnover. Biochem. Biophys.<br />

Res. Commun., 287, 995-1002.<br />

[27] Godiksen S, Selzer-Plon J, Pedersen ED, Abell K, Rasmussen HB, Szabo R, Bugge TH, & Vogel<br />

LK (2008) Hepatocyte growth factor activator inhibitor-1 has a complex subcellular itinerary.<br />

Biochem. J., 413, 251-259.<br />

[28] Oberst MD, Williams CA, Dickson RB, Johnson MD, & Lin CY (2003) The activation of<br />

matriptase requires its noncatalytic domains, serine protease domain, and its cognate<br />

inhibitor. J. Biol. Chem., 278, 26773-26779.<br />

[29] Lee MS, Tseng IC, Wang Y, Kiyomiya K, Johnson MD, Dickson RB, & Lin CY (2007)<br />

Autoactivation of matriptase in vitro: requirement for biomembrane and LDL receptor<br />

domain. Am. J. Physiol Cell Physiol, 293, C95-105.<br />

[30] Lee MS, Kiyomiya K, Benaud C, Dickson RB, & Lin CY (2005) Simultaneous activation and<br />

hepatocyte growth factor activator inhibitor 1-mediated inhibition of matriptase induced at<br />

activation foci in human mammary epithelial cells. Am. J. Physiol Cell Physiol, 288, C932-<br />

C941.<br />

[31] Benaud C, Dickson RB, & Lin CY (2001) Regulation of the activity of matriptase on epithelial<br />

cell surfaces by a blood-derived factor. Eur. J. Biochem., 268, 1439-1447.<br />

[32] Lopez-Otin C & Matrisian LM (2007) Emerging roles of proteases in tumour suppression. Nat.<br />

Rev. Cancer, 7, 800-808.<br />

[33] Hedstrom L (2002) Serine protease mechanism and specificity. Chem. Rev., 102, 4501-4524.<br />

[34] Bugge TH, Antalis TM, & Wu Q (2009) Type II transmembrane serine proteases. J. Biol. Chem.,<br />

284, 23177-23181.<br />

[35] Netzel-Arnett S, Hooper JD, Szabo R, Madison EL, Quigley JP, Bugge TH, & Antalis TM (2003)<br />

Membrane anchored serine proteases: a rapidly expanding group of cell surface proteolytic<br />

enzymes with potential roles in cancer. Cancer Metastasis Rev., 22, 237-258.<br />

[36] Hooper JD, Clements JA, Quigley JP, & Antalis TM (2001) Type II transmembrane serine<br />

proteases. Insights into an emerging class of cell surface proteolytic enzymes. J. Biol. Chem.,<br />

276, 857-860.<br />

[37] Szabo R & Bugge TH (2008) Type II transmembrane serine proteases in development and<br />

disease. Int. J. Biochem. Cell Biol., 40, 1297-1316.<br />

[38] Kim MG, Chen C, Lyu MS, Cho EG, Park D, Kozak C, & Schwartz RH (1999) Cloning and<br />

chromosomal mapping of a gene isolated from thymic stromal cells encoding a new mouse<br />

type II membrane serine protease, epithin, containing four LDL receptor modules and two<br />

CUB domains. Immunogenetics, 49, 420-428.<br />

[39] Lin CY, Anders J, Johnson M, Sang QA, & Dickson RB (1999) Molecular cloning of cDNA for<br />

matriptase, a matrix-degrading serine protease with trypsin-like activity. J. Biol. Chem., 274,<br />

18231-18236.<br />

93


[40] Takeuchi T, Shuman MA, & Craik CS (1999) Reverse biochemistry: use of macromolecular<br />

protease inhibitors to dissect complex biological processes and identify a membrane-type<br />

serine protease in epithelial cancer and normal tissue. Proc. Natl. Acad. Sci. U. S. A, 96,<br />

11054-11061.<br />

[41] Cao J, Cai X, Zheng L, Geng L, Shi Z, Pao CC, & Zheng S (1997) Characterization of colorectalcancer-related<br />

cDNA clones obtained by subtractive hybridization screening. J. Cancer Res.<br />

Clin. Oncol., 123, 447-451.<br />

[42] Underwood LJ, Tanimoto H, Wang Y, Shigemasa K, Parmley TH, & O'Brien TJ (1999) Cloning of<br />

tumor-associated differentially expressed gene-14, a novel serine protease overexpressed by<br />

ovarian carcinoma. Cancer Res., 59, 4435-4439.<br />

[43] Shi YE, Torri J, Yieh L, Wellstein A, Lippman ME, & Dickson RB (1993) Identification and<br />

characterization of a novel matrix-degrading protease from hormone-dependent human<br />

breast cancer cells. Cancer Res., 53, 1409-1415.<br />

[44] Lin CY, Anders J, Johnson M, & Dickson RB (1999) Purification and characterization of a<br />

complex containing matriptase and a Kunitz-type serine protease inhibitor from human milk.<br />

J. Biol. Chem., 274, 18237-18242.<br />

[45] Velasco G, Cal S, Quesada V, Sanchez LM, & Lopez-Otin C (2002) Matriptase-2, a membranebound<br />

mosaic serine proteinase predominantly expressed in human liver and showing<br />

degrading activity against extracellular matrix proteins. J. Biol. Chem., 277, 37637-37646.<br />

[46] Szabo R, Netzel-Arnett S, Hobson JP, Antalis TM, & Bugge TH (2005) Matriptase-3 is a novel<br />

phylogenetically preserved membrane-anchored serine protease with broad serpin<br />

reactivity. Biochem. J., 390, 231-242.<br />

[47] Ge W, Hu H, Ding K, Sun L, & Zheng S (2006) Protein interaction analysis of ST14 domains and<br />

their point and deletion mutants. J. Biol. Chem., 281, 7406-7412.<br />

[48] Antalis TM, Buzza MS, Hodge KM, Hooper JD, & Netzel-Arnett S (2010) The cutting edge:<br />

membrane-anchored serine protease activities in the pericellular microenvironment.<br />

Biochem. J., 428, 325-346.<br />

[49] Hobson JP, Netzel-Arnett S, Szabo R, Rehault SM, Church FC, Strickland DK, Lawrence DA,<br />

Antalis TM, & Bugge TH (2004) Mouse DESC1 is located within a cluster of seven DESC1-like<br />

genes and encodes a type II transmembrane serine protease that forms serpin inhibitory<br />

complexes. J. Biol. Chem., 279, 46981-46994.<br />

[50] Beliveau F, Desilets A, & Leduc R (2009) Probing the substrate specificities of matriptase,<br />

matriptase-2, hepsin and DESC1 with internally quenched fluorescent peptides. FEBS J., 276,<br />

2213-2226.<br />

[51] Takeuchi T, Harris JL, Huang W, Yan KW, Coughlin SR, & Craik CS (2000) Cellular localization<br />

of membrane-type serine protease 1 and identification of protease-activated receptor-2 and<br />

single-chain urokinase-type plasminogen activator as substrates. J. Biol. Chem., 275, 26333-<br />

26342.<br />

94


[52] Friedrich R, Fuentes-Prior P, Ong E, Coombs G, Hunter M, Oehler R, Pierson D, Gonzalez R,<br />

Huber R, Bode W, & Madison EL (2002) Catalytic domain structures of MT-SP1/matriptase, a<br />

matrix-degrading transmembrane serine proteinase. J. Biol. Chem., 277, 2160-2168.<br />

[53] Lee SL, Dickson RB, & Lin CY (2000) Activation of hepatocyte growth factor and<br />

urokinase/plasminogen activator by matriptase, an epithelial membrane serine protease. J.<br />

Biol. Chem., 275, 36720-36725.<br />

[54] Owen KA, Qiu D, Alves J, Schumacher AM, Kilpatrick LM, Li J, Harris JL, & Ellis V (2010)<br />

Pericellular activation of hepatocyte growth factor by the transmembrane serine proteases<br />

matriptase and hepsin, but not by the membrane-associated protease uPA. Biochem. J., 426,<br />

219-228.<br />

[55] Bhatt AS, Welm A, Farady CJ, Vasquez M, Wilson K, & Craik CS (2007) Coordinate expression<br />

and functional profiling identify an extracellular proteolytic signaling pathway. Proc. Natl.<br />

Acad. Sci. U. S. A, 104, 5771-5776.<br />

[56] Kilpatrick LM, Harris RL, Owen KA, Bass R, Ghorayeb C, Bar-Or A, & Ellis V (2006) Initiation of<br />

plasminogen activation on the surface of monocytes expressing the type II transmembrane<br />

serine protease matriptase. Blood, 108, 2616-2623.<br />

[57] Camerer E, Barker A, Duong DN, Ganesan R, Kataoka H, Cornelissen I, Darragh MR, Hussain A,<br />

Zheng YW, Srinivasan Y, Brown C, Xu SM, Regard JB, Lin CY, Craik CS, Kirchhofer D, & Coughlin<br />

SR (2010) Local protease signaling contributes to neural tube closure in the mouse embryo.<br />

Dev. Cell, 18, 25-38.<br />

[58] Jin X, Yagi M, Akiyama N, Hirosaki T, Higashi S, Lin CY, Dickson RB, Kitamura H, & Miyazaki K<br />

(2006) Matriptase activates stromelysin (MMP-3) and promotes tumor growth and<br />

angiogenesis. Cancer Sci., 97, 1327-1334.<br />

[59] Sales KU, Masedunskas A, Bey AL, Rasmussen AL, Weigert R, List K, Szabo R, Overbeek PA, &<br />

Bugge TH (2010) Matriptase initiates activation of epidermal pro-kallikrein and disease onset<br />

in a mouse model of Netherton syndrome. Nat. Genet., 42, 676-683.<br />

[60] Bhatt AS, Erdjument-Bromage H, Tempst P, Craik CS, & Moasser MM (2005) Adhesion<br />

signaling by a novel mitotic substrate of src kinases. Oncogene, 24, 5333-5343.<br />

[61] Ahmed S, Jin X, Yagi M, Yasuda C, Sato Y, Higashi S, Lin CY, Dickson RB, & Miyazaki K (2006)<br />

Identification of membrane-bound serine proteinase matriptase as processing enzyme of<br />

insulin-like growth factor binding protein-related protein-1 (IGFBPrP1/angiomodulin/mac25).<br />

FEBS J., 273, 615-627.<br />

[62] Darragh MR, Bhatt AS, & Craik CS (2008) MT-SP1 proteolysis and regulation of cellmicroenvironment<br />

interactions. Front Biosci., 13, 528-539.<br />

[63] Hurst NJ, Jr., Najy AJ, Ustach CV, Movilla L, & Kim HR (2012) Platelet-derived growth factor-C<br />

(PDGF-C) activation by serine proteases: implications for breast cancer progression. Biochem.<br />

J., 441, 909-918.<br />

[64] Ustach CV, Huang W, Conley-Lacomb MK, Lin CY, Che M, Abrams J, & Kim HR (2010) A Novel<br />

Signaling Axis of Matriptase/PDGF-D/{beta}-PDGFR in Human Prostate Cancer. Cancer Res.,<br />

70, 9631-9640.<br />

95


[65] Chen M, Chen LM, Lin CY, & Chai KX (2009) Hepsin activates prostasin and cleaves the<br />

extracellular domain of the epidermal growth factor receptor. Mol. Cell Biochem..<br />

[66] Chen M, Chen LM, Lin CY, & Chai KX (2008) The epidermal growth factor receptor (EGFR) is<br />

proteolytically modified by the Matriptase-Prostasin serine protease cascade in cultured<br />

epithelial cells. Biochim. Biophys. Acta, 1783, 896-903.<br />

[67] Tripathi M, Potdar AA, Yamashita H, Weidow B, Cummings PT, Kirchhofer D, & Quaranta V<br />

(2011) Laminin-332 cleavage by matriptase alters motility parameters of prostate cancer<br />

cells. Prostate, 71, 184-196.<br />

[68] Kim C, Cho Y, Kang CH, Kim MG, Lee H, Cho EG, & Park D (2005) Filamin is essential for<br />

shedding of the transmembrane serine protease, epithin. EMBO Rep., 6, 1045-1051.<br />

[69] Gal P & Zavodszky P (1998) Structure and function of the serine-protease subcomponents of<br />

C1: protein engineering studies. Immunobiology, 199, 317-326.<br />

[70] Janciauskiene S, Nita I, Subramaniyam D, Li Q, Lancaster JR, Jr., & Matalon S (2008) Alpha1-<br />

antitrypsin inhibits the activity of the matriptase catalytic domain in vitro. Am. J. Respir. Cell<br />

Mol. Biol., 39, 631-637.<br />

[71] Tseng IC, Chou FP, Su SF, Oberst M, Madayiputhiya N, Lee MS, Wang JK, Sloane DE, Johnson<br />

M, & Lin CY (2008) Purification from human milk of matriptase complexes with secreted<br />

serpins: mechanism for inhibition of matriptase other than HAI-1. Am. J. Physiol Cell Physiol,<br />

295, C423-C431.<br />

[72] Chou FP, Xu H, Lee MS, Chen YW, Richards OX, Swanson R, Olson ST, Johnson MD, & Lin CY<br />

(2011) Matriptase is inhibited by extravascular antithrombin in epithelial cells but not in most<br />

carcinoma cells. Am. J. Physiol Cell Physiol, 301, C1093-C1103.<br />

[73] Yamasaki Y, Satomi S, Murai N, Tsuzuki S, & Fushiki T (2003) Inhibition of membrane-type<br />

serine protease 1/matriptase by natural and synthetic protease inhibitors. J. Nutr. Sci.<br />

Vitaminol. (Tokyo), 49, 27-32.<br />

[74] Myerburg MM, McKenna EE, Luke CJ, Frizzell RA, Kleyman TR, & Pilewski JM (2008) Prostasin<br />

expression is regulated by airway surface liquid volume and is increased in cystic fibrosis. Am.<br />

J. Physiol Lung Cell Mol. Physiol, 294, L932-L941.<br />

[75] Kataoka H, Itoh H, & Koono M (2002) Emerging multifunctional aspects of cellular serine<br />

proteinase inhibitors in tumor progression and tissue regeneration. Pathol. Int., 52, 89-102.<br />

[76] Laskowski M, Jr. & Kato I (1980) Protein inhibitors of proteinases. Annu. Rev. Biochem., 49,<br />

593-626.<br />

[77] Goder V & Spiess M (2001) Topogenesis of membrane proteins: determinants and dynamics.<br />

FEBS Lett., 504, 87-93.<br />

[78] Wang JK, Lee MS, Tseng IC, Chou FP, Chen YW, Fulton A, Lee HS, Chen CJ, Johnson MD, & Lin<br />

CY (2009) Polarized epithelial cells secrete matriptase as a consequence of zymogen<br />

activation and HAI-1-mediated inhibition. Am. J. Physiol Cell Physiol, 297, C459-C470.<br />

96


[79] Szabo R, Hobson JP, List K, Molinolo A, Lin CY, & Bugge TH (2008) Potent inhibition and global<br />

co-localization implicate the transmembrane Kunitz-type serine protease inhibitor<br />

hepatocyte growth factor activator inhibitor-2 in the regulation of epithelial matriptase<br />

activity. J. Biol. Chem., 283, 29495-29504.<br />

[80] Shimomura T, Denda K, Kitamura A, Kawaguchi T, Kito M, Kondo J, Kagaya S, Qin L, Takata H,<br />

Miyazawa K, & Kitamura N (1997) Hepatocyte growth factor activator inhibitor, a novel<br />

Kunitz-type serine protease inhibitor. J. Biol. Chem., 272, 6370-6376.<br />

[81] Kawaguchi T, Qin L, Shimomura T, Kondo J, Matsumoto K, Denda K, & Kitamura N (1997)<br />

Purification and cloning of hepatocyte growth factor activator inhibitor type 2, a Kunitz-type<br />

serine protease inhibitor. J. Biol. Chem., 272, 27558-27564.<br />

[82] Akiyama Y, Nagai M, Komaki W, Marutsuka K, Asada Y, & Kataoka H (2006) Expression of<br />

hepatocyte growth factor activator inhibitor type 1 in endothelial cells. Hum. Cell, 19, 91-97.<br />

[83] Szabo R, Hobson JP, Christoph K, Kosa P, List K, & Bugge TH (2009) Regulation of cell surface<br />

protease matriptase by HAI2 is essential for placental development, neural tube closure and<br />

embryonic survival in mice. Development, 136, 2653-2663.<br />

[84] Denda K, Shimomura T, Kawaguchi T, Miyazawa K, & Kitamura N (2002) Functional<br />

characterization of Kunitz domains in hepatocyte growth factor activator inhibitor type 1. J.<br />

Biol. Chem., 277, 14053-14059.<br />

[85] Girard TJ, Warren LA, Novotny WF, Likert KM, Brown SG, Miletich JP, & Broze GJ, Jr. (1989)<br />

Functional significance of the Kunitz-type inhibitory domains of lipoprotein-associated<br />

coagulation inhibitor. Nature, 338, 518-520.<br />

[86] Denda K, Shimomura T, Kawaguchi T, Miyazawa K, & Kitamura N (2002) Functional<br />

characterization of Kunitz domains in hepatocyte growth factor activator inhibitor type 1. J.<br />

Biol. Chem., 277, 14053-14059.<br />

[87] Qin L, Denda K, Shimomura T, Kawaguchi T, & Kitamura N (1998) Functional characterization<br />

of Kunitz domains in hepatocyte growth factor activator inhibitor type 2. FEBS Lett., 436,<br />

111-114.<br />

[88] Kojima K, Tsuzuki S, Fushiki T, & Inouye K (2008) Roles of functional and structural domains<br />

of hepatocyte growth factor activator inhibitor type 1 in the inhibition of matriptase. J. Biol.<br />

Chem., 283, 2478-2487.<br />

[89] Levitin F, Stern O, Weiss M, Gil-Henn C, Ziv R, Prokocimer Z, Smorodinsky NI, Rubinstein DB,<br />

& Wreschner DH (2005) The MUC1 SEA module is a self-cleaving domain. J. Biol. Chem., 280,<br />

33374-33386.<br />

[90] Cho EG, Kim MG, Kim C, Kim SR, Seong IS, Chung C, Schwartz RH, & Park D (2001) N-terminal<br />

processing is essential for release of epithin, a mouse type II membrane serine protease. J.<br />

Biol. Chem., 276, 44581-44589.<br />

[91] Miyake Y, Yasumoto M, Tsuzuki S, Fushiki T, & Inouye K (2009) Activation of a membranebound<br />

serine protease matriptase on the cell surface. J. Biochem., 146, 273-282.<br />

97


[92] Bode W, Schwager P, & Huber R (1978) The transition of bovine trypsinogen to a trypsin-like<br />

state upon strong ligand binding. The refined crystal structures of the bovine trypsinogenpancreatic<br />

trypsin inhibitor complex and of its ternary complex with Ile-Val at 1.9 A<br />

resolution. J. Mol. Biol., 118, 99-112.<br />

[93] Qiu D, Owen K, Gray K, Bass R, & Ellis V (2007) Roles and regulation of membrane-associated<br />

serine proteases. Biochem. Soc. Trans., 35, 583-587.<br />

[94] Thielens NM, Bersch B, Hernandez JF, & Arlaud GJ (1999) Structure and functions of the<br />

interaction domains of C1r and C1s: keystones of the architecture of the C1 complex.<br />

Immunopharmacology, 42, 3-13.<br />

[95] Miyake Y, Tsuzuki S, Mochida S, Fushiki T, & Inouye K (2010) The role of asparagine-linked<br />

glycosylation site on the catalytic domain of matriptase in its zymogen activation. Biochim.<br />

Biophys. Acta, 1804, 156-165.<br />

[96] Oberst MD, Chen LY, Kiyomiya K, Williams CA, Lee MS, Johnson MD, Dickson RB, & Lin CY<br />

(2005) HAI-1 regulates activation and expression of matriptase, a membrane-bound serine<br />

protease. Am. J. Physiol Cell Physiol, 289, C462-C470.<br />

[97] Kiyomiya K, Lee MS, Tseng IC, Zuo H, Barndt RJ, Johnson MD, Dickson RB, & Lin CY (2006)<br />

Matriptase activation and shedding with HAI-1 is induced by steroid sex hormones in human<br />

prostate cancer cells, but not in breast cancer cells. Am. J. Physiol Cell Physiol, 291, C40-C49.<br />

[98] Tseng IC, Xu H, Chou FP, Li G, Vazzano AP, Kao JP, Johnson MD, & Lin CY (2010) Matriptase<br />

activation, an early cellular response to acidosis. J. Biol. Chem., 285, 3261-3270.<br />

[99] Inouye K, Yasumoto M, Tsuzuki S, Mochida S, & Fushiki T (2010) The optimal activity of a<br />

pseudozymogen form of recombinant matriptase under the mildly acidic pH and low ionic<br />

strength conditions. J. Biochem., 147, 485-492.<br />

[100] Cho EG, Schwartz RH, & Kim MG (2005) Shedding of membrane epithin is blocked without<br />

LDLRA4 and its protease activation site. Biochem. Biophys. Res. Commun., 327, 328-334.<br />

[101] Yu JX, Chao L, & Chao J (1994) Prostasin is a novel human serine proteinase from seminal<br />

fluid. Purification, tissue distribution, and localization in prostate gland. J. Biol. Chem., 269,<br />

18843-18848.<br />

[102] Yu JX, Chao L, & Chao J (1995) Molecular cloning, tissue-specific expression, and cellular<br />

localization of human prostasin mRNA. J. Biol. Chem., 270, 13483-13489.<br />

[103] Vallet V, Chraibi A, Gaeggeler HP, Horisberger JD, & Rossier BC (1997) An epithelial serine<br />

protease activates the amiloride-sensitive sodium channel. Nature, 389, 607-610.<br />

[104] Adachi M, Kitamura K, Miyoshi T, Narikiyo T, Iwashita K, Shiraishi N, Nonoguchi H, & Tomita K<br />

(2001) Activation of epithelial sodium channels by prostasin in Xenopus oocytes. J. Am. Soc.<br />

Nephrol., 12, 1114-1121.<br />

[105] Chen LM, Skinner ML, Kauffman SW, Chao J, Chao L, Thaler CD, & Chai KX (2001) Prostasin is<br />

a glycosylphosphatidylinositol-anchored active serine protease. J. Biol. Chem., 276, 21434-<br />

21442.<br />

98


[106] Rickert KW, Kelley P, Byrne NJ, Diehl RE, Hall DL, Montalvo AM, Reid JC, Shipman JM, Thomas<br />

BW, Munshi SK, Darke PL, & Su HP (2008) Structure of human prostasin, a target for the<br />

regulation of hypertension. J. Biol. Chem., 283, 34864-34872.<br />

[107] Shipway A, Danahay H, Williams JA, Tully DC, Backes BJ, & Harris JL (2004) Biochemical<br />

characterization of prostasin, a channel activating protease. Biochem. Biophys. Res.<br />

Commun., 324, 953-963.<br />

[108] Fan B, Wu TD, Li W, & Kirchhofer D (2005) Identification of hepatocyte growth factor<br />

activator inhibitor-1B as a potential physiological inhibitor of prostasin. J. Biol. Chem., 280,<br />

34513-34520.<br />

[109] Chen LM, Zhang X, & Chai KX (2004) Regulation of prostasin expression and function in the<br />

prostate. Prostate, 59, 1-12.<br />

[110] Scharschmidt TC, List K, Grice EA, Szabo R, Renaud G, Lee CC, Wolfsberg TG, Bugge TH, &<br />

Segre JA (2009) Matriptase-deficient mice exhibit ichthyotic skin with a selective shift in skin<br />

microbiota. J. Invest Dermatol., 129, 2435-2442.<br />

[111] Suzuki Y, Nomura J, Koyama J, & Horii I (1994) The role of proteases in stratum corneum:<br />

involvement in stratum corneum desquamation. Arch. Dermatol. Res., 286, 249-253.<br />

[112] Komatsu N, Takata M, Otsuki N, Ohka R, Amano O, Takehara K, & Saijoh K (2002) Elevated<br />

stratum corneum hydrolytic activity in Netherton syndrome suggests an inhibitory regulation<br />

of desquamation by SPINK5-derived peptides. J. Invest Dermatol., 118, 436-443.<br />

[113] Chavanas S, Bodemer C, Rochat A, Hamel-Teillac D, Ali M, Irvine AD, Bonafe JL, Wilkinson J,<br />

Taieb A, Barrandon Y, Harper JI, de PY, & Hovnanian A (2000) Mutations in SPINK5, encoding<br />

a serine protease inhibitor, cause Netherton syndrome. Nat. Genet., 25, 141-142.<br />

[114] Hewett DR, Simons AL, Mangan NE, Jolin HE, Green SM, Fallon PG, & McKenzie AN (2005)<br />

Lethal, neonatal ichthyosis with increased proteolytic processing of filaggrin in a mouse<br />

model of Netherton syndrome. Hum. Mol. Genet., 14, 335-346.<br />

[115] Chen CJ, Wu BY, Tsao PI, Chen CY, Wu MH, Chan YL, Lee HS, Johnson MD, Eckert RL, Chen<br />

YW, Chou F, Wang JK, & Lin CY (2010) Increased matriptase zymogen activation in<br />

inflammatory skin disorders. Am. J. Physiol Cell Physiol.<br />

[116] Leyvraz C, Charles RP, Rubera I, Guitard M, Rotman S, Breiden B, Sandhoff K, & Hummler E<br />

(2005) The epidermal barrier function is dependent on the serine protease CAP1/Prss8. J. Cell<br />

Biol., 170, 487-496.<br />

[117] Mildner M, Ballaun C, Stichenwirth M, Bauer R, Gmeiner R, Buchberger M, Mlitz V, &<br />

Tschachler E (2006) Gene silencing in a human organotypic skin model. Biochem. Biophys.<br />

Res. Commun., 348, 76-82.<br />

[118] Miller GS & List K (2012) The matriptase-prostasin proteolytic cascade in epithelial<br />

development and pathology. Cell Tissue Res..<br />

[119] Ovaere P, Lippens S, Vandenabeele P, & Declercq W (2009) The emerging roles of serine<br />

protease cascades in the epidermis. Trends Biochem. Sci., 34, 453-463.<br />

99


[120] Vuagniaux G, Vallet V, Jaeger NF, Hummler E, & Rossier BC (2002) Synergistic activation of<br />

ENaC by three membrane-bound channel-activating serine proteases (mCAP1, mCAP2, and<br />

mCAP3) and serum- and glucocorticoid-regulated kinase (Sgk1) in Xenopus Oocytes. J. Gen.<br />

Physiol, 120, 191-201.<br />

[121] Mauro T, Guitard M, Behne M, Oda Y, Crumrine D, Komuves L, Rassner U, Elias PM, &<br />

Hummler E (2002) The ENaC channel is required for normal epidermal differentiation. J.<br />

Invest Dermatol., 118, 589-594.<br />

[122] Planes C, Randrianarison NH, Charles RP, Frateschi S, Cluzeaud F, Vuagniaux G, Soler P, Clerici<br />

C, Rossier BC, & Hummler E (2010) ENaC-mediated alveolar fluid clearance and lung fluid<br />

balance depend on the channel-activating protease 1. EMBO Mol. Med., 2, 26-37.<br />

[123] Andreasen D, Vuagniaux G, Fowler-Jaeger N, Hummler E, & Rossier BC (2006) Activation of<br />

epithelial sodium channels by mouse channel activating proteases (mCAP) expressed in<br />

Xenopus oocytes requires catalytic activity of mCAP3 and mCAP2 but not mCAP1. J. Am. Soc.<br />

Nephrol., 17, 968-976.<br />

[124] Vallet V, Pfister C, Loffing J, & Rossier BC (2002) Cell-surface expression of the channel<br />

activating protease xCAP-1 is required for activation of ENaC in the Xenopus oocyte. J. Am.<br />

Soc. Nephrol., 13, 588-594.<br />

[125] Netzel-Arnett S, Buzza MS, Shea-Donohue T, Desilets A, Leduc R, Fasano A, Bugge TH, &<br />

Antalis TM (2011) Matriptase protects against experimental colitis and promotes intestinal<br />

barrier recovery. Inflamm. Bowel. Dis..<br />

[126] Fan B, Brennan J, Grant D, Peale F, Rangell L, & Kirchhofer D (2007) Hepatocyte growth factor<br />

activator inhibitor-1 (HAI-1) is essential for the integrity of basement membranes in the<br />

developing placental labyrinth. Dev. Biol., 303, 222-230.<br />

[127] Tanaka H, Nagaike K, Takeda N, Itoh H, Kohama K, Fukushima T, Miyata S, Uchiyama S,<br />

Uchinokura S, Shimomura T, Miyazawa K, Kitamura N, Yamada G, & Kataoka H (2005)<br />

Hepatocyte growth factor activator inhibitor type 1 (HAI-1) is required for branching<br />

morphogenesis in the chorioallantoic placenta. Mol. Cell Biol., 25, 5687-5698.<br />

[128] Szabo R, Molinolo A, List K, & Bugge TH (2007) Matriptase inhibition by hepatocyte growth<br />

factor activator inhibitor-1 is essential for placental development. Oncogene, 26, 1546-1556.<br />

[129] Nagaike K, Kawaguchi M, Takeda N, Fukushima T, Sawaguchi A, Kohama K, Setoyama M, &<br />

Kataoka H (2008) Defect of hepatocyte growth factor activator inhibitor type 1/serine<br />

protease inhibitor, Kunitz type 1 (Hai-1/Spint1) leads to ichthyosis-like condition and<br />

abnormal hair development in mice. Am. J. Pathol., 173, 1464-1475.<br />

[130] Heinz-Erian P, Muller T, Krabichler B, Schranz M, Becker C, Ruschendorf F, Nurnberg P,<br />

Rossier B, Vujic M, Booth IW, Holmberg C, Wijmenga C, Grigelioniene G, Kneepkens CM,<br />

Rosipal S, Mistrik M, Kappler M, Michaud L, Doczy LC, Siu VM, Krantz M, Zoller H, Utermann<br />

G, & Janecke AR (2009) Mutations in SPINT2 cause a syndromic form of congenital sodium<br />

diarrhea. Am. J. Hum. Genet., 84, 188-196.<br />

[131] Hanahan D & Weinberg RA (2000) The hallmarks of cancer. Cell, 100, 57-70.<br />

100


[132] Tanimoto H, Underwood LJ, Wang Y, Shigemasa K, Parmley TH, & O'Brien TJ (2001) Ovarian<br />

tumor cells express a transmembrane serine protease: a potential candidate for early<br />

diagnosis and therapeutic intervention. Tumour. Biol., 22, 104-114.<br />

[133] Santin AD, Cane' S, Bellone S, Bignotti E, Palmieri M, De Las Casas LE, Anfossi S, Roman JJ,<br />

O'Brien T, & Pecorelli S (2003) The novel serine protease tumor-associated differentially<br />

expressed gene-15 (matriptase/MT-SP1) is highly overexpressed in cervical carcinoma.<br />

Cancer, 98, 1898-1904.<br />

[134] Tsai WC, Chu CH, Yu CP, Sheu LF, Chen A, Chiang H, & Jin JS (2008) Matriptase and survivin<br />

expression associated with tumor progression and malignant potential in breast cancer of<br />

Chinese women: tissue microarray analysis of immunostaining scores with clinicopathological<br />

parameters. Dis. Markers, 24, 89-99.<br />

[135] Kang JY, Dolled-Filhart M, Ocal IT, Singh B, Lin CY, Dickson RB, Rimm DL, & Camp RL (2003)<br />

Tissue microarray analysis of hepatocyte growth factor/Met pathway components reveals a<br />

role for Met, matriptase, and hepatocyte growth factor activator inhibitor 1 in the<br />

progression of node-negative breast cancer. Cancer Res., 63, 1101-1105.<br />

[136] Jin JS, Cheng TF, Tsai WC, Sheu LF, Chiang H, & Yu CP (2007) Expression of the serine<br />

protease, matriptase, in breast ductal carcinoma of Chinese women: correlation with<br />

clinicopathological parameters. Histol. Histopathol., 22, 305-309.<br />

[137] Riddick AC, Shukla CJ, Pennington CJ, Bass R, Nuttall RK, Hogan A, Sethia KK, Ellis V, Collins<br />

AT, Maitland NJ, Ball RY, & Edwards DR (2005) Identification of degradome components<br />

associated with prostate cancer progression by expression analysis of human prostatic<br />

tissues. Br. J. Cancer, 92, 2171-2180.<br />

[138] Saleem M, Adhami VM, Zhong W, Longley BJ, Lin CY, Dickson RB, Reagan-Shaw S, Jarrard DF,<br />

& Mukhtar H (2006) A novel biomarker for staging human prostate adenocarcinoma:<br />

overexpression of matriptase with concomitant loss of its inhibitor, hepatocyte growth factor<br />

activator inhibitor-1. Cancer Epidemiol. Biomarkers Prev., 15, 217-227.<br />

[139] Lee JW, Yong SS, Choi JJ, Lee SJ, Kim BG, Park CS, Lee JH, Lin CY, Dickson RB, & Bae DS (2005)<br />

Increased expression of matriptase is associated with histopathologic grades of cervical<br />

neoplasia. Hum. Pathol., 36, 626-633.<br />

[140] Zeng L, Cao J, & Zhang X (2005) Expression of serine protease SNC19/matriptase and its<br />

inhibitor hepatocyte growth factor activator inhibitor type 1 in normal and malignant tissues<br />

of gastrointestinal tract. World J. Gastroenterol., 11, 6202-6207.<br />

[141] Vogel LK, Saebo M, Skjelbred CF, Abell K, Pedersen ED, Vogel U, & Kure EH (2006) The ratio of<br />

Matriptase/HAI-1 mRNA is higher in colorectal cancer adenomas and carcinomas than<br />

corresponding tissue from control individuals. BMC. Cancer, 6, 176.<br />

[142] Welman A, Sproul D, Mullen P, Muir M, Kinnaird AR, Harrison DJ, Faratian D, Brunton VG, &<br />

Frame MC (2012) Diversity of matriptase expression level and function in breast cancer.<br />

PLoS. One., 7, e34182.<br />

[143] Zhang Y, Cai X, Schlegelberger B, & Zheng S (1998) Assignment1 of human putative tumor<br />

suppressor genes ST13 (alias SNC6) and ST14 (alias SNC19) to human chromosome bands<br />

22q13 and 11q24-->q25 by in situ hybridization. Cytogenet. Cell Genet., 83, 56-57.<br />

101


[144] Kawaguchi M, Takeda N, Hoshiko S, Yorita K, Baba T, Sawaguchi A, Nezu Y, Yoshikawa T,<br />

Fukushima T, & Kataoka H (2011) Membrane-bound serine protease inhibitor HAI-1 is<br />

required for maintenance of intestinal epithelial integrity. Am. J. Pathol., 179, 1815-1826.<br />

[145] Oberst MD, Johnson MD, Dickson RB, Lin CY, Singh B, Stewart M, Williams A, al-Nafussi A,<br />

Smyth JF, Gabra H, & Sellar GC (2002) Expression of the serine protease matriptase and its<br />

inhibitor HAI-1 in epithelial ovarian cancer: correlation with clinical outcome and tumor<br />

clinicopathological parameters. Clin. Cancer Res., 8, 1101-1107.<br />

[146] Baba T, Kawaguchi M, Fukushima T, Sato Y, Orikawa H, Yorita K, Tanaka H, Lin CY, Sakoda S, &<br />

Kataoka H (2012) Loss of membrane-bound serine protease inhibitor HAI-1 induces oral<br />

squamous cell carcinoma cells' invasiveness. J. Pathol..<br />

[147] Sambuy Y, de A, I, Ranaldi G, Scarino ML, Stammati A, & Zucco F (2005) The Caco-2 cell line as<br />

a model of the intestinal barrier: influence of cell and culture-related factors on Caco-2 cell<br />

functional characteristics. Cell Biol. Toxicol., 21, 1-26.<br />

[148] Misfeldt DS, Hamamoto ST, & Pitelka DR (1976) Transepithelial transport in cell culture. Proc.<br />

Natl. Acad. Sci. U. S. A, 73, 1212-1216.<br />

[149] Nimishakavi S, Besprozvannaya M, Raymond WW, Craik CS, Gruenert DC, & Caughey GH<br />

(2012) Activity and Inhibition of Prostasin and Matriptase on Apical and Basolateral Surfaces<br />

of Human Airway Epithelial Cells. Am. J. Physiol Lung Cell Mol. Physiol.<br />

[150] Buzza MS, Netzel-Arnett S, Shea-Donohue T, Zhao A, Lin CY, List K, Szabo R, Fasano A, Bugge<br />

TH, & Antalis TM (2010) Membrane-anchored serine protease matriptase regulates epithelial<br />

barrier formation and permeability in the intestine. Proc. Natl. Acad. Sci. U. S. A.<br />

[151] Thermo Scientific. Immunoprecipitation Methodology. 2012.<br />

Ref Type: Online Source<br />

[152] Thermo Scientific. Thermo Scientific Avidin-Biotin Technical Handbook. 2012.<br />

Ref Type: Online Source<br />

[153] Hoffman M & Monroe DM, III (2001) A cell-based model of hemostasis. Thromb. Haemost.,<br />

85, 958-965.<br />

[154] Yu IS, Chen HJ, Lee YS, Huang PH, Lin SR, Tsai TW, & Lin SW (2000) Mice deficient in hepsin, a<br />

serine protease, exhibit normal embryogenesis and unchanged hepatocyte regeneration<br />

ability. Thromb. Haemost., 84, 865-870.<br />

[155] Wu Q, Yu D, Post J, Halks-Miller M, Sadler JE, & Morser J (1998) Generation and<br />

characterization of mice deficient in hepsin, a hepatic transmembrane serine protease. J.<br />

Clin. Invest, 101, 321-326.<br />

[156] Bruns JB, Carattino MD, Sheng S, Maarouf AB, Weisz OA, Pilewski JM, Hughey RP, & Kleyman<br />

TR (2007) Epithelial Na+ channels are fully activated by furin- and prostasin-dependent<br />

release of an inhibitory peptide from the gamma-subunit. J. Biol. Chem., 282, 6153-6160.<br />

[157] Svenningsen P, Uhrenholt TR, Palarasah Y, Skjodt K, Jensen BL, & Skott O (2009) Prostasindependent<br />

activation of epithelial Na+ channels by low plasmin concentrations. Am. J.<br />

Physiol Regul. Integr. Comp Physiol, 297, R1733-R1741.<br />

102


[158] Powers JC, Asgian JL, Ekici OD, & James KE (2002) Irreversible inhibitors of serine, cysteine,<br />

and threonine proteases. Chem. Rev., 102, 4639-4750.<br />

[159] Schoellmann G & Shaw E (1963) Direct evidence for the presence of histidine in the active<br />

center of chymotrypsin. Biochemistry, 2, 252-255.<br />

[160] Anderson PJ & Bock PE (2001) Biotin derivatives of D-Phe-Pro-Arg-CH2Cl for active-sitespecific<br />

labeling of thrombin and other serine proteinases. Anal. Biochem., 296, 254-261.<br />

[161] Ranby M, Bergsdorf N, & Nilsson T (1982) Enzymatic properties of the one- and two-chain<br />

form of tissue plasminogen activator. Thromb. Res., 27, 175-183.<br />

[162] Boose JA, Kuismanen E, Gerard R, Sambrook J, & Gething MJ (1989) The single-chain form of<br />

tissue-type plasminogen activator has catalytic activity: studies with a mutant enzyme that<br />

lacks the cleavage site. Biochemistry, 28, 635-643.<br />

[163] Tate KM, Higgins DL, Holmes WE, Winkler ME, Heyneker HL, & Vehar GA (1987) Functional<br />

role of proteolytic cleavage at arginine-275 of human tissue plasminogen activator as<br />

assessed by site-directed mutagenesis. Biochemistry, 26, 338-343.<br />

[164] Tachias K & Madison EL (1995) Variants of tissue-type plasminogen activator which display<br />

substantially enhanced stimulation by fibrin. J. Biol. Chem., 270, 18319-18322.<br />

[165] Madison EL, Kobe A, Gething MJ, Sambrook JF, & Goldsmith EJ (1993) Converting tissue<br />

plasminogen activator to a zymogen: a regulatory triad of Asp-His-Ser. Science, 262, 419-421.<br />

[166] Darragh MR, Schneider EL, Lou J, Phojanakong PJ, Farady CJ, Marks JD, Hann BC, & Craik CS<br />

(2010) Tumor detection by imaging proteolytic activity. Cancer Res., 70, 1505-1512.<br />

[167] Galkin AV, Mullen L, Fox WD, Brown J, Duncan D, Moreno O, Madison EL, & Agus DB (2004)<br />

CVS-3983, a selective matriptase inhibitor, suppresses the growth of androgen independent<br />

prostate tumor xenografts. Prostate, 61, 228-235.<br />

[168] Farady CJ, Sun J, Darragh MR, Miller SM, & Craik CS (2007) The mechanism of inhibition of<br />

antibody-based inhibitors of membrane-type serine protease 1 (MT-SP1). J. Mol. Biol., 369,<br />

1041-1051.<br />

[169] Enyedy IJ, Lee SL, Kuo AH, Dickson RB, Lin CY, & Wang S (2001) Structure-based approach for<br />

the discovery of bis-benzamidines as novel inhibitors of matriptase. J. Med. Chem., 44, 1349-<br />

1355.<br />

[170] Long YQ, Lee SL, Lin CY, Enyedy IJ, Wang S, Li P, Dickson RB, & Roller PP (2001) Synthesis and<br />

evaluation of the sunflower derived trypsin inhibitor as a potent inhibitor of the type II<br />

transmembrane serine protease, matriptase. Bioorg. Med. Chem. Lett., 11, 2515-2519.<br />

[171] Li P, Jiang S, Lee SL, Lin CY, Johnson MD, Dickson RB, Michejda CJ, & Roller PP (2007) Design<br />

and synthesis of novel and potent inhibitors of the type II transmembrane serine protease,<br />

matriptase, based upon the sunflower trypsin inhibitor-1. J. Med. Chem., 50, 5976-5983.<br />

[172] Sun J, Pons J, & Craik CS (2003) Potent and selective inhibition of membrane-type serine<br />

protease 1 by human single-chain antibodies. Biochemistry, 42, 892-900.<br />

103


[173] Napp J, Dullin C, Muller F, Uhland K, Petri JB, van de Locht A, Steinmetzer T, & Alves F (2010)<br />

Time-domain in vivo near infrared fluorescence imaging for evaluation of matriptase as a<br />

potential target for the development of novel, inhibitor-based tumor therapies. Int. J. Cancer,<br />

127, 1958-1974.<br />

[174] Steinmetzer T, Schweinitz A, Sturzebecher A, Donnecke D, Uhland K, Schuster O, Steinmetzer<br />

P, Muller F, Friedrich R, Than ME, Bode W, & Sturzebecher J (2006) Secondary amides of<br />

sulfonylated 3-amidinophenylalanine. New potent and selective inhibitors of matriptase. J.<br />

Med. Chem., 49, 4116-4126.<br />

[175] Farady CJ, Egea PF, Schneider EL, Darragh MR, & Craik CS (2008) Structure of an Fab-protease<br />

complex reveals a highly specific non-canonical mechanism of inhibition. J. Mol. Biol., 380,<br />

351-360.<br />

104


Supplementary I<br />

Hepatocyte growth factor activator inhibitor-1 has a complex subcellular itinerary<br />

Sine Godiksen 1 , Joanna Selzer‐Plon 1 , Esben D. K. Pedersen 1 , Kathrine Abell 1 , Hanne Borger<br />

Rasmussen 2 , Roman Szabo 3 , Thomas H. Bugge 3 , and Lotte K. Vogel 1<br />

1 Department of Cellular and Molecular Medicine, University of Copenhagen, Copenhagen,<br />

Denmark.<br />

2 Department of Biomedical Science, University of Copenhagen, Copenhagen, Denmark<br />

3 Proteases and Tissue Remodeling Unit, National <strong>Institut</strong>e of Dental and Craniofacial Research,<br />

National <strong>Institut</strong>es of Health, Bethesda, USA<br />

Published in Biochemical Journal, July 2008.<br />

105

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!