extracellular matrix proteins as drivers of inflammation ...20schi%f8ler%20schultz.pdf · mekanisme...

120
UNIVERSITY OF COPENHAGEN FACULTY OF SCIENCE Industrial PhD Thesis Heidi Schiøler Schultz Extracellular matrix proteins as drivers of inflammation in rheumatoid arthritis Academic advisor: Martin W. Berchtold Submitted: 13/03/2015

Upload: lamdiep

Post on 19-Aug-2019

214 views

Category:

Documents


0 download

TRANSCRIPT

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

F A C U L T Y O F S C I E N C E

Industrial PhD Thesis

Heidi Schiøler Schultz

Extracellular matrix proteins as drivers of

inflammation in rheumatoid arthritis Thesis subtitle

Academic advisor: Martin W. Berchtold

Submitted: 13/03/2015

2

Extracellular matrix proteins as drivers of

inflammation in rheumatoid arthritis

Industrial PhD Thesis

Heidi Schiøler Schultz

Department of Biology, Section for Cell and Developmental Biology, Faculty of Science,

University of Copenhagen.

Copenhagen, Denmark

&

Department of Immuobiology, AID, Novo Nordisk A/S

Måløv, Denmark

2015

”It is good to have an end to journey towards, but it is the journey that matters in the end”

– Ursula K. Le Guin

3

Extracellular matrix proteins as drivers of inflammation in rheumatoid arthritis

The PhD project was financially supported by the Danish Ministry of Science, Technology and

Innovation and R&D Academic Relations, Novo Nordisk A/S

This thesis was submitted to the Graduate School of the Faculty of Science, University of

Copenhagen, March 2015

Principal supervisor

Professor Martin W. Berchtold

Department of Cell and Developmental Biology, Faculty of Science,

University of Copenhagen, Denmark

Co-supervisors

Principal Scientist, PhD, MD, Svetlana Panina

Department of Immunobiology

Novo Nordisk A/S, Måløv, Denmark

Senior Scientist, PhD Christine Brender Read

Department of Autoimmune Disease Research

Novo Nordisk A/S, Måløv, Denmark

Principal Scientist, PhD, Olle Björkdahl

Department of Immunogenicity Assessment

Novo Nordisk A/S, Måløv, Denmark

Assessment committee

Senior Research Fellow, PhD, Kalle Søderstrøm

NDORMS, Botnar Research Center

Oxford University, United Kingdom

4

Research Scientist, PhD, Anders Aspberg

Department of Clinical Sciences

Lund University, Lund, Sweden

Section Head, DSc, Niels Behrendt (chairman)

Finsen Laboratory, Rigshospitalet,

BRIC, University of Copenhagen, Denmark

Cover:

The drawing shows pathways of differentiation of monocyte-derived cells within the joints of

patients with rheumatoid arthritis. The drawings are made by S. Panina and kindly given to me.

PhD Thesis 2015 © Heidi Schiøler Schultz

5

List of publications

Papers included in this PhD Thesis

I. Heidi S. Schultz, Louise M. Nitze,, Louise H. Zeuthen, Pernille Keller, Albrecht Gruhler,

Jesper Pass, Jianhe Chen, Li Guo, Andrew J. Fleetwood, John A. Hamilton, Martin W.

Berchtold and Svetlana Panina. (2015) Collagen Induces Maturation of Human

Monocyte-Derived Dendritic Cells by Signaling through Osteoclast-Associated. J.

Immunology. 194 (epub, doi 10.4049/jimmunol.1402800)

II. Heidi S. Schultz, Li Guo, Pernille Keller, Andrew J. Fleetwood, Wei Guo, John A.

Hamilton, Olle Björkdahl, Martin W. Berchtold and Svetlana Panina. OSCAR-collagen

signaling mediates activation of monocytes. To be submitted to J. Immunology March,

2015.

Poster abstracts

III. Extracellular matrix proteins may control tissue-specific maturation of dendritic

cells: implications for therapy of Rheumatoid Arthritis. Heidi S. Schultz, Louise M.

Nitze, Louise H. Zeuthen, Pernille Keller, Li Guo and Svetlana Panina. Poster presented

at the PhD Day: Building research through dissemination, Department of Biology,

University of Copenhagen, 2013 Nov 1-14th

.

Relevant papers not included in the PhD

IV. Fleetwood, A. J., A. Achuthan, H. Schultz, A. Nansen, K. Almholt, P. Usher, and J. A.

Hamilton. (2014). Urokinase plasminogen activator is a central regulator of

macrophage three-dimensional invasion, matrix degradation, and adhesion. J.

Immunology. 192: 3540-3547.

6

Abstract

Rheumatoid arthritis (RA) is a chronic systemic autoimmune disease that primarily affects the

peripheral diarthroidal joints. The joint inflammation in RA is characterized by hyperplasia of the

synovium and pannus formation as well as by the presence of pro-inflammatory cytokines and

chemokines which causes a massive influx of immune cells into the synovial tissue (ST). A

hallmark of RA pathogenesis is the enhanced formation and activation of osteoclasts (OC). The

progressive inflammation leads to remodeling of the extracellular matrix (ECM), and ultimately

results in destruction of cartilage and bone in the joint. RA can to some degree be suppressed by

non-biologics disease-modifying anti-rheumatic drugs (nbDMARDS) or biologics (bDMARDS)

such as e.g. TNF-α inhibitors. However, the efficacy of the current drugs is limited as they are only

suppressing inflammation, but are not able to halt joint destruction completely. Identification of

factors that regulate inflammation and bone/cartilage loss is crucial for development of a new

generation of drugs to treat RA.

The OSteoClast-Associated Receptor (OSCAR) was proposed as a new target for

treatment of RA. OSCAR has been reported to be up-regulated in the synovium of RA patients and

on the circulating monocytes where its expression level has been shown to positively correlate with

the disease activity. In mice, OSCAR expression is limited to OC. It was proposed that OSCAR

provides a co-stimulatory signaling in RANK-dependent osteoclastogenesis. In humans, OSCAR is

expressed in OC, DCs, monocytes, macrophages and neutrophils. We and others have identified the

ligands of OSCAR to be collagen type I (ColI) and –type II (ColII). In RA, excessive degradation of

ECM by enzymes secreted by activated immune cells, fibroblasts and OC leads to collagens

exposure and makes them available for interaction with the immune cells. Thus, we hypothesize

that at such pathological conditions ColI/II serve as naturally occurring activators of the OSCAR

signaling.

In the present study, we have investigated the functional outcome of the OSCAR-

collagen interaction in human myeloid cells. We have observed that OSCAR engagement by ColI/II

induced activation and maturation of DCs characterized by up-regulation of cell surface markers

and secretion of pro-inflammatory cytokines. These collagen-matured DCs (Col-DCs) were

efficient drivers of naïve T cell proliferation. The T cells expanded by Col-DCs secreted cytokines

with no clear T-cell polarization pattern. Global RNA profiling revealed that multiple pro-

inflammatory mediators, including cytokines and cytokine receptors, components of the stable

7

immune synapse (namely, CD40, CD86, CD80, ICAM-1), as well as components of TNF- and TLR

signaling, are transcriptional targets of OSCAR in DCs.

The functional role of the OSCAR-collagen interaction was further addressed in

monocytes and macrophages. We have shown that the OSCAR-ColII signaling supports survival of

monocytes under conditions of growth factor withdrawal. Moreover, ColII stimulated release of

pro-inflammatory cytokines by monocytes from healthy donors, and this release could be

completely blocked by an antagonistic anti-OSCAR mAb. Similarly, mononuclear cells from

synovial fluid of RA patients cultured on ColII secreted TNF-α and IL-8 in an OSCAR-dependent

manner. Global profiling of gene expression in monocytes showed that components of multiple

signaling pathways relevant for the RA pathogenesis are transcriptional targets of OSCAR.

Surprisingly, OSCAR engagement by ColII in macrophages, which were differentiated in vitro from

monocytes of healthy donors, did not lead to secretion of cytokines. Microarray analysis did not

reveal any gene regulation by the OSCAR-ColII signaling in macrophages. However, degradation

of collagen-rich matrix, gelatin, by macrophages was significantly abolished by an antagonistic

anti-OSCAR mAb, indicating that OSCAR modulates the activity of proteolytic enzymes on a post-

transcriptional level.

Taken together, our findings indicate the existence of a novel pathway mediating a

sustained inflammation in the RA joints where collagens become exposed during tissue remodeling

and can thus activate the OSCAR signaling in monocyte-derived cells. We hypothesize that the

OSCAR-collagen pathway can potentially contribute to the RA pathogenesis on multiple levels and

may thus represents a new target for therapeutic intervention.

8

Sammendrag (Danish summary)

Kronisk leddegigt (reumatoid artrit) er en systemisk, autoimmun sygdom, der hovedsageligt

påvirker de perifere led. Karakteristisk for betændelsestilstanden i leddegigt er en øget celledeling

af synovialvævet og tilstedeværelsen af pro-inflammatoriske cytokiner og chemokiner, som øger

tilstrømning af immunceller til synovialleddet. Et af kendetegnene ved leddegigt er den øgede

dannelse og aktivering af de knoglenedbrydende osteoklaster. Den progressive betændelsestilstand

medfører en omstrukturering af den ekstracellulære matrix, som fører til ødelæggelse af knogle og

brusk i leddet. Leddegigt kan til en vis grad hæmmes af sygdomsmodificerende lægemidler, som

methotrexate, eller biologiske lægemidler såsom TNF-α inhibitors. Desværre er effekten af de

nuværende behandlingsformer begrænset, da det i de fleste tilfælde kun er betændelsestilstanden der

mindskes og ikke nedbrydningen af knogler og brusk. Det er derfor vigtigt for udviklingen af ny

behandling til gigtpatienter, at identificere nye faktorer der kan regulere både betændelsestilstanden

og knogle/brusk nedbrydelsen.

Som en ny behandlingsform af leddegigtspatienter er receptoren OSteoClast-

Associated Receptor (OSCAR) blevet foreslået. Studier har vist, at der er et øget niveau af OSCAR

i leddene hos leddegigtpatienter, og at forekomsten af OSCAR på monocytter i blodet korrelerer

positivt med sygdomsaktiviteten. I mus er OSCAR kun udtrykt på osteoklaster, hvor OSCAR ses at

have en co-stimulatorisk rolle i RANK-medieret osteoklast-dannelse. I mennesket er OSCAR

udtrykt på både osteoklaster, dendritiske celler, monocytter, makrofager og neutrofile granulocytter.

Liganden til OSCAR har vi og andre identificeret til at være Collagen type I og -II.

I leddegigt udskiller aktiveret immunceller, fibroblaster og osteoklaster høje mængder

enzym, som nedbryder den ekstracellulære matrix, hvilket blotter collagen og muliggør interaktion

mellem collagenen og immuncellerne i leddet. Under de patologiske forhold i forbindelse med

leddegigt, antages det derfor, at collagen type I og -II vil fungere som naturlige ligander for OSCAR

signalering.

Vi har i dette studie undersøgt den funktionelle betydning af interaktionen mellem

OSCAR og collagen i celler af myeloid oprindelse. Vi observerede at interaktionen mellem OSCAR

og collagen øgede aktiveringen og modningen af dendritiske celler, hvilket var karakteriseret ved

opregulering af overflademarkører og udskillelse af pro-inflammatoriske cytokiner. Endvidere så vi,

at de dendritiske celler, som var modnet med collagen, effektivt stimulerede T celle deling, men

disse T celler udviste ikke nogen klar cytokin polarisering. Genanalyse af de collagen stimulerede

9

dendritiske celler viste at flere betændelsesmarkører blev opreguleret, såsom cytokiner,

cytokinereceptorer, dele af den stabiliserende immunologiske synapse (herunder CD40, CD86,

CD80, ICAM-1) og dele af TNF- og TLR signalering.

Derudover blev den funktionelle rolle af OSCAR-collagen interaktionen også

undersøgt i monocytter og makrofager. I fravær af overlevelsesfaktorer kunne vi i monocytter

konstatere, at OSCAR-collagen interaktionen forlængede deres overlevelse. Derudover stimulerede

collagen udskillelsen af flere pro-inflammatoriske cytokiner i monocytter fra raske kontroller,

hvilket vi kunne hæmme ved at tilsætte et antagonistisk anti-OSCAR antistof. Global genanalyse

viste at flere transkriptionelle mål for OSCAR signaleringen i monocytter er involveret i

patogenesen for leddegigt. Vi noterede ydermere at celler fra synovialvæsken fra patienter med

leddegigt udskilte TNF-α og IL-8 når de blev stimuleret med collagen, hvilket var afhængigt af

OSCAR. Overraskende nok så vi ikke cytokin-udskillelse ved collagen-stimulering af monocyte-

genereret makrofager og tilsvarende så vi ingen effekt på gen niveau. Imidlertid konstaterede vi,

ved at hæmme OSCAR signaleringen i makrofager, at nedbrydelsen af collagen-rig matrix blev

mindsket, hvilket kunne indikere at OSCAR kan modulere enzym aktiviteten post-transkriptionelt.

Tilsammen indikerer vores fund at OSCAR-collagen interaktionen er en vigtig

mekanisme der bidrager til betændelsestilstanden i leddet hos patienter med leddegigt, hvor

collagen bliver blottet på grund af strukturelle ændringer i ledvævet og derfor kan interagere med

OSCAR i monocyt-genererede myeloide celler. Vi mener derfor at OSCAR-collagen interaktionen

kan bidrage til øget forståelse af patogenesen i leddegigt på flere niveauer og at mekanismen

ligeledes kan bruges som nyt, effektivt mål for udvikling af behandling.

10

Preface

This thesis was written in order to obtain the PhD degree from the Faculty of Science, University of

Copenhagen. The thesis is the result of three years experimental work at the Biopharmaceutical

Research Unit, Novo Nordisk A/S, Måløv, Denmark, and Department of Biology, University of

Copenhagen, Denmark. The thesis was carried out in the period between April 2012 and March

2015. Professor Martin W. Berchtold was my internal supervisor at Department of Cell and

Developmental Biology, Faculty of Science, University of Copenhagen. My main external

supervisor was Principal Scientist Svetlana Panina, Department of Immunobiology,

Biopharmaceutical Research Unit, Novo Nordisk A/S. A part of the PhD studies was carried out at

the University of Melbourne, Australia, under the supervision of Professor John Hamilton and

Research Fellow Andrew Fleetwood.

The PhD project was proposed by Novo Nordisk A/S in connection with the drug discovery

program for treatment of rheumatoid arthritis. During my PhD study, several changes took place in

the R&D research focus and the company structure. In December 2013 Novo Nordisk A/S decided

to discontinue the OSCAR project. The OSCAR project was closed due to challenges connected

with restricted expression of OSCAR in rodents, making it difficult to conduct in vivo studies using

animal disease models. I took a decision to continue with the OSCAR project as described in the

PhD proposal with adjustments to keep it in line with company activities. In March 2014, Novo

Nordisk re-organized their whole inflammation area, resulting in closure of Dept. of

Immunobiology. I continued my PhD in Department of Cellular Pharmacology, under the

supervision of Scientist and Team leader Christine Brender Read. In September 2014, the company

took a strategic decision to discontinue all activities within autoimmune inflammation disease

(AID) research and the whole function area was closed down. I therefore continued my PhD in

Immunogenicity Assessment, Non-Clinical development, being supervised by Principal Scientist

Olle Björkdahl.

The PhD thesis consists of an introduction summarizing the current knowledge about the interface

between the bone biology and inflammation. It outlines the pathogenesis of rheumatoid arthritis,

focusing on the contribution of immune cells, cytokines and extracellular matrix, as well as the

immunoreceptor OSCAR. The results are arranged in tree sections, based on two manuscripts

11

(results part I and part II) and a section written in a paper format, called results part III,

summarizing unpublished data. Paper I describes the role of OSCAR in dendritic cells and it was

recently published in Journal of Immunology. Paper II focuses on the OSCAR function in

monocytes and is currently being prepared for submission to Journal of Immunology. The third

section shows preliminary data on the OSCAR biology in macrophages. The methodologies used to

conduct the studies are described in the two papers and in part III of the results. Finally, the results

are discussed in a general Discussion section based on the two papers and my unpublished data on

the OSCAR biology in the context of rheumatoid arthritis.

________________________________

Copenhagen, March 2015

Heidi Schultz

12

Acknowledgements

It has been three educating and challenging years during my PhD studies, which I could not have

done without help from my supervisors, colleagues, collaborators, friends and family. It has been

truly rewarding to be a part of the daily routine at the departments of Immunobiology, Cellular

Pharmacology and Immunogenicity Assessment.

Many colleagues have contributed to the work in this thesis and I would like to direct my gratitude

to several people.

Special thanks to my first supervisor Svetlana Panina. Svetlana was the Biology Coordinator of the

OSCAR project and proposed my PhD project. Svetlana has a deep scientific knowledge and a great

passion for science. I would like to thank Svetlana for her expertise, interesting discussions and

encouragement when needed. Svetlana has my deepest respect for still finding time and making the

effort to guide and help me in the last stretch of my PhD, even though she has not been my official

supervisor since June 2014. For that I am very grateful.

Professor Martin Berchtold has been my internal university supervisor. He has engaged me in his

group and challenged me every time he could, encouraging me in my work. Your advice has always

been of great help.

I would also like to thank my supervisor Christine Brender Read, who, even though it was for a

short period of time, welcomed me into her little but very competent team.

Last but not least, thanks to Olle Björkdahl. Olle was so kind to be my supervisor in the challenging

period after Inflammation research was closed down. He made sure that I had what I needed to

finish my PhD. I would like to thank Olle for engaging in me and my project, asking questions and

guiding me through the last part. Special thanks for always offering your help willingly, and for

rapid and constructive corrections of my writings.

13

Special thanks to Professor John A. Hamilton, Andrew Fleetwood and colleagues at the Arthritis

and Inflammation Research Centre, Department of Medicine, Royal Melbourne Hospital, Australia,

for an amazing stay. It has been rewarding in many ways both scientifically and personally.

I wish to acknowledge everyone who contributed to my scientific work, in particular Pernille Keller

for her help with micro array analysis, as well as Lise W. F. Jensen, Gou Li, Lotte Bentzon and

Sofie Hedlund Møller. Thanks to all my colleagues, especially in Inflammation Biology for

practical help with laboratory techniques and encouraging talks. Thanks to the PhD student Mette

Dandanell Nielsen, Kasper Vadstrup and master student Pernille Damgaard Petersen for daily chats

and problem solving discussions. I thank Suzi Høgh Madsen for letting me borrow some of her own

figures for my introduction (Fig. 2 and 5).

An acknowledgement to the Danish Ministry of Science, Technology and Innovation and R&D

Academic relations at Novo Nordisk A/S for financial support.

I want to express my gratitude to the assessment committee for evaluating this thesis.

Finally, I wish to direct my thankfulness to friends and family. Special thanks to Mie and Palle S.

Schultz, Ninna S. Hansen and Andreas E. Clemmensen for their unconditional love and support.

The strength gained from such love is tremendous. Andreas, you are the best part of our little team.

14

Abbreviations

(m)Ab: (Monoclonal) Antibody

Ag: Antigen

APC: Allophycocyanin

APC: Antigen-presenting cell

CCL: CC Chemokine ligand

CCLR: CC chemokine receptor

CD: Cluster of differentiation

Col I/II: Collagen type 1 or 2

CIA: Collagen-induced arthritis

CXCL: CXC chemokine ligand

CXCR: CXC chemokine receptor

Cy7: Cyanin-7

DAP12: DNAX activation protein of 12 kDa

(i)DC: (immature) Dendritic cell

DDR: Discoidin-domain receptor

DMARDs: Disease-modifying anti-rheumatic drugs

DMSO: Dimethyl sulfoxide

ECM: Extracellular matrix

EDTA: Ethylene diamine tetra acetic acid

FACS: Fluorescence activated cell sorter

FcR: Fc receptor common gamma chain

FCS: Fetal calf serum

FITC: Fluorescein isothiocyanate

FSC: Forward scatter

GM-CSF: Granulocyte-macrophage colony-stimulating factor

GPVI: Glycoprotein VI

HLA: Human leukocyte antigen

HSA: Human serum albumin

HSC: Hematopoietic stem cell

ICAM-1: Intracellular cell adhesion molecule-1

Ig: Immunoglobulin

IL: Interleukin

INF: Interferon gamma

ITAM: Immunoreceptor tyrosine-based activation motif

ITIM: Immunoreceptor tyrosine-based inhibition motif

LAIR-1: Leukocyte-associated Ig-like receptor-1

LPS: Lipopolysaccharide

MAPK: Mitogen-activated protein kinases

MCP-1: Monocyte chemotactic protein-1

15

M-CSF: Macrophage colony-stimulating factor

MHC: Major Histocompatibility complex

MIP: Macrophage inflammatory protein

MMP: Matrix metalloproteinase

MNC: Mononuclear cells

Mo: Monocyte

NFATc1: Nuclear factor of activated T-cells

NFB: Nuclear factor kappa B

NSAID: Non-steroidal anti-inflammatory drug

OB: Osteoblasts

OC: Osteoclast

OPG: Osteoprotegerin

(s)OSCAR: (soluble) Osteoclast-associated receptor

PAI: Plasminogen activator inhibitor

PBMC: Peripheral blood mononuclear cells

PBS: Phosphate buffer saline

PE: Phycoerythrin

PerCP: Peridinin chlorophyll protein

Plg: Plasminogen

P/S: Penicillin/Streptomycin

RA: Rheumatoid arthritis

RANK: Receptor activator of nuclear factor kappa B

RANKL: RANK ligand

RANTES: Regulated and normal T cell expressed and secreted

RBC: Red blood cell

Rh: Recombinant human

RNA: Ribonucleic acid

rpm: Revolutions per minute

SD: Standard deviation

SSC: Side scatter

STAT: Signal transducer and activator of transcription

TCR: T cell receptor

TGF-β: Transforming growth factor beta

Th: T helper

TLR: Toll-like receptors

TNF-: Tumor necrosis factor alpha

TNFR: TNF receptor

TNP: Trinitrophenyl

TRAP: Tartrate-resistant acid phosphatase

TREM: Triggering receptor expressed on myeloid cells

uPA: Urokinase-type plasminogen activator

VCAM-1: Vascular cell adhesion molecule-1

16

Table of contents

1. INTRODUCTION........................................................................................................... 18

1.1 Osteoimmunology ..................................................................................................................................................... 18 1.1.1 Immune cells ....................................................................................................................................................... 18 1.1.2 Cells regulating bone homeostasis ...................................................................................................................... 20

1.2 Rheumatoid arthritis ................................................................................................................................................ 23 1.2.1 Cellular contribution to rheumatoid arthritis ....................................................................................................... 24 1.2.2 Cytokines and chemokines in rheumatoid arthritis ............................................................................................. 26 1.2.3 Rheumatoid arthritis treatment ............................................................................................................................ 27

1.3 Extracellular matrix ................................................................................................................................................. 28 1.3.1 Extracellular matrix of bone and cartilage .......................................................................................................... 28 1.3.2 Extracellular matrix turnover .............................................................................................................................. 30 1.3.3 Collagen receptors .............................................................................................................................................. 32

1.4 Osteoclast-associated receptor (OSCAR) ............................................................................................................... 35 1.4.1 Murine OSCAR .................................................................................................................................................. 35 1.4.2 Human OSCAR .................................................................................................................................................. 36 1.4.3 OSCAR function ................................................................................................................................................. 36 1.4.4 OSCAR ligand .................................................................................................................................................... 37 1.4.5 Regulation of OSCAR ........................................................................................................................................ 38 1.4.6 OSCAR in disease ............................................................................................................................................... 38

2. RATIONALE AND AIM ................................................................................................. 40

3. RESULTS ....................................................................................................................... 41

3.1 OSCAR function in dendritic cells .......................................................................................................................... 42

3.2 OSCAR function in monocytes ................................................................................................................................ 54 3.2.1 Abstract ............................................................................................................................................................... 55 3.2.2 Introduction ......................................................................................................................................................... 56 3.2.3 Materials and methods ........................................................................................................................................ 58 3.2.4 Results ................................................................................................................................................................. 62 3.2.5 Discussion ........................................................................................................................................................... 69

3.3 OSCAR function in macrophages ........................................................................................................................... 73 3.3.1 Introduction ......................................................................................................................................................... 73 3.3.2 Material and methods .......................................................................................................................................... 74 3.3.3 Results ................................................................................................................................................................. 78 3.3.4 Discussion ........................................................................................................................................................... 84

4. DISCUSSION ................................................................................................................. 87

5. CONCLUSION ............................................................................................................... 94

6. PERSPECTIVES ............................................................................................................ 95

6.1 OSCAR project ......................................................................................................................................................... 95

17

6.2 Atherosclerosis .......................................................................................................................................................... 96

6.3 Limitations to the OSCAR project .......................................................................................................................... 96

7. REFERENCES ............................................................................................................... 98

8. APPENDIX ................................................................................................................... 114

8.1 Appendix I ............................................................................................................................................................... 114

8.2 Appendix II ............................................................................................................................................................. 116

8.3 Co-author contributions ......................................................................................................................................... 117

18

1. Introduction

1.1 Osteoimmunology

Osteoimmunology is a new field of research that deals with the role of the immune system in bone

remodeling1. Immune cells and bone cells are closely interconnected sharing signaling molecules

and site of origin, namely the bone marrow. Attention was drawn to this field due to the observation

that bone destruction can be caused by an abnormal activation of the immune system. Osteoclast-

mediated bone loss has been observed in various autoimmune diseases, like rheumatoid arthritis and

diabetes mellitus2. On the contrary, it has also been reported that bone cells influence immune

cells3.

1.1.1 Immune cells

The cellular component of the immune system consists of two main types, the myeloid and the

lymphoid cells. Both lineages derive from the hematopoietic stem cells (HSCs). The myeloid cells

represent an integral part of the natural immune system and participate in the first-line response

against infectious agents. The myeloid cells are phagocytic cells that scavenge debris generated by

physiological or pathological processes4.

Fig. 1. Development of the cells of the immune system. Immune cells originate from the hematopoietic stem cells in

the bone marrow. Hematopoietic stem cells are capable of generating progenitor cells that will later give rise to different

types of immune cells. There are two main types of progenitors, the myeloid progenitors and the lymphoid progenitors.

The figure was produced by Schultz, H. S. using Servier Medical art figures as a template.

19

Monocytes derive from bone marrow progenitors, circulate in the blood and are capable of

differentiating into macrophages, dendritic cells and osteoclasts depending on signals from the

extracellular environment. Monocytes can function as antigen-presenting cells (APCs) and can

secrete various inflammatory factors5.

Macrophages are heterogenic tissue-resident cells. They have high proteolytic and

catabolic activities contributing to their ability to scavenge and digest pathogens, dead cells and

cellular debris. Macrophages have the capability to present antigens, though they do it less efficient

compared to other APCs like dendritic cells (DCs) and B-cells. At a steady-state, they are

considered to be anti-inflammatory, maintaining organ homeostasis5. Under inflammatory

conditions, they become activated and take part in host response to pathogens6.

DCs are specialized APCs which play an important role in processing and presentation

of antigens (Ag) to T cells. DCs secrete cytokines important for both innate and adaptive immunity.

At present, four main DCs types are categorized being: conventional DCs (cDCs), plasmacytoid

DCs (pDCs), Langerhans cells and monocyte-derived DCs. These subsets constitute a

heterogeneous population, classified further by their anatomical location, phenotype, functional

properties and migratory capability7;8

. Most human DCs are differentiated from blood monocytes or

monocyte precursors to immature DCs (iDCs) under the control of granulocyte-macrophage colony-

stimulating factor (GM-CSF) and Interleukin (IL)-49. DCs migrate from the bone marrow to

peripheral tissues, where they reside in an immature state, awaiting antigen encounter. iDCs can

become mature in response to several stimuli, like IL-1, TNF-, LPS, CD40-CD40L interaction or

by exposure to extracellular matrix (ECM) components, such as collagen10

. The activity of DCs can

be regulated by Toll-like receptors (TLR), as well as activating and inhibiting immunoreceptors,

which transduce signals trough immunoreceptor tyrosine-based activating motifs (ITAM) and

through immunoreceptor tyrosine-based inhibiting motifs (ITIM), respectively11;12

. When

encountering an Ag the DCs process the Ag and display the peptides via their MHC complex.

While maturing, the DCs migrate to the T cells areas of secondary lymphoid organs, where they

present Ag to T cells. The outcome - the immune priming or tolerance - depends on the maturation

stage of the DCs13-15

. DCs are important for orchestrating the adaptive immune response and may

also contribute to the development of autoimmunity when dysregulated16

.

Important for the initiation of the adaptive immune response is T cell activation. T

cells are activated by interaction between the T cell receptor (TCR) on the T cell and the major

histocompatibility complex (MHC)-Ag peptide complex on the APCs. The TCR has a low affinity

20

for the MHC molecule, but the interaction is strengthened by the formation of a specialized contact,

an immunological synapse. The immunological synapse is characterized by a specific pattern of

receptor segregation with a central cluster of TCR surrounded by co-stimulatory receptors and a

ring of integrin adhesion molecules17

. For T cells to be activated two signals are needed. The first

signal is the peptide-MHC:TCR complex, stabilized by either CD4 or CD8 co-receptor. The second

signal is provided by the co-stimulatory molecules such as CD80 and CD86 on the APCs that

interacts with CD28 on the T cells. Without co-stimulation T cells become anergic, a mechanism to

protect against autoimmunity. The interaction between the T cell and the APC is enhanced by

adhesion molecules, like lymphocyte function-associated antigen 1 (LFA-1) expressed on T cells

and intercellular adhesion molecule 1 (ICAM-1) present on APC18

. The outcome of T cell

activation is dependent on the state of APC maturation, e.g. number of MHC-peptide complexes as

well as duration of TCR engagement. The activation generally results in T cell proliferation and

polarization, and expression of multiple cytokines including IL-2, IFNγ and TNF-19;20

.

1.1.2 Cells regulating bone homeostasis

Bone contains several types of cells that are important for its development, maintenance and repair.

In adults, bone is constantly being remodeled, enabling the tissue to regenerate damages, adapt to

changing stress and to control calcium homeostasis. The main cell types facilitating bone

remodeling are the specialized bone-resorbing cells, osteoclasts (OCs) and the bone-synthesizing

cells, osteoblasts (OBs)2.

Bone remodeling (Fig. 2A) is initiated by osteocytes21;22

. Osteocytes are DC-like cells

trapped in the bone matrix. Osteocytes act as mechanosensors in bone tissue and activate other

osteocytes and bone lining cells when microdamage occurs. This signaling results in attraction of

OC precursors and initiation of osteoclastogenesis. OCs originates from cells of the myeloid

lineage, which fuse into large multinucleated OCs. Essential for OCs differentiation, activity and

survival are macrophage-colony stimulating factor (M-CSF) and receptor activator of nuclear

factor-kB ligand (RANKL). M-CSF induces early differentiation, promotes proliferation and

survival of OCs precursors. RANKL is a member of the TNF superfamily and signals through its

receptor, RANK, which is expressed on a variety of cells including OCs. RANK-RANKL

interaction induces the final differentiation of OCs and their bone-resorbing activity23-27

. M-CSF

and RANKL are believed to be sufficient for inducing OC differentiation in culture28

. However, co-

stimulatory signals mediated by ITAM-harboring receptors have been proven to be important co-

stimulatory signaling component in osteoclastogenesis and are essential for bone remodeling29-32

.

21

At sites of active bone resorption, OCs forms a specialized cell membrane, namely the

ruffled border. The ruffled cell border seals and degrades the underlying bone in the resorption pit

through the release of hydrogen ions (lowering of pH) and intracellular secretory vesicles

containing enzymes like tartrate-resistant acid phosphatase (TRAP) and protease cathepsin-K (Fig.

2B)26

. When OCs have completed their job, they undergo apoptosis, and OB precursors are

attracted to the site, where they differentiate into mature OBs. The OBs synthesize collagen-rich

organic matrix and provide optimal conditions for matrix mineralization. OBs turn into osteocytes

after being surrounded by bone matrix, thus stopping matrix synthesis and mineralization2;33

.

Fig. 2. A. The bone remodeling cycle. A microcrack or old

bone causes osteocytic apoptosis, which induces RANKL

expression by bone lining cells. This attracts monocytes to the

damaged bone area, and osteoclastogenesis starts. The

osteoclasts will start to resorb bone, which is followed by

formation of new bone matrix by the osteoblasts. Osteoblasts

that are trapped in the remodeled matrix become osteocytes,

whereas the rest either die or become flattened osteoblast lining

cells. B. The resorbing osteoclast seals tightly to the bone

surface to create a closed local environment known as the

resorption lacuna. Bone resorption takes place, by lowering the

pH (secretion of H+) and by releasing proteases by

exocytosis26;27

. Figure borrowed from Suzi H. Madsen.

22

Modulation of osteoclastogenesis

OCs function is mainly regulated by TNF-, TNFR and TNF-like proteins such as osteoprotegerin

(OPG), RANK and RANKL. Additionally, hormones, cytokines and humoral factors regulate bone

density. TNF-, IL-1, IL-6 and IL-17 increase bone-resorption by locally enhancing RANKL

expression34

, which together with IL-1β promotes survival of OCs26

.

The RANK signaling pathway is negatively regulated by OPG. OPG blocks OCs

formation in vitro and bone resorption in vivo35

. OPG functions by blocking RANKL binding to

RANK. Accessibility of RANKL and OPG is coordinated to regulate bone resorption by controlling

the activation state of RANK on OCs26

. Some cytokines like IFN- and IL-4 have an inhibitory

effect on osteoclastogenesis36-38

.

Trans-differentiation

During in vivo inflammatory bone conditions, considerable de novo osteoclastogenesis is observed.

It has been investigated whether cell types other than those from the monocyte/macrophage lineage

can generate multinucleated giant cell with bone resorption capacity. Several recent in vitro studies

have reported that both human (monocyte-derived) and mouse (bone marrow-derived) DCs can

develop into functional OCs when cultured in the presence of M-CSF, RANKL and bone-like

matrix9;39-41

. This process is greatly enhanced by the microenvironment in the joint, involving IL-

1, TNF- and the extracellular matrix hyaluronic acid9;40

. The osteoclastogenic potential is

restricted to iDCs, which upon cell maturation is lost, even after addition of RANKL42

.

Additionally, in mice, it has been observed that CD11c+

DCs upon interaction with CD4+

T cells

develop into functional OCs in response to microbial or protein Ags in a RANKL/RANK-

dependent manner39

. Taken together, these findings indicate a high plasticity of DCs and suggest

that DCs may directly contribute to osteoclastogenesis.

23

1.2 Rheumatoid arthritis

Rheumatoid arthritis (RA) is a chronic systemic autoimmune inflammatory disease that targets the

articular cartilage and bone at the joint margins, as well as periarticular and subchondral bone,

leading to joint destruction43

. It is characterized by the presence of autoantibodies, synovial

inflammation, pannus formation, cartilage damage and bone erosion. RA affects up to 1% of the

population, women more often than men, and is associated with significant morbidity and

mortality44

. The molecular mechanisms involved in RA pathogenesis remain unclear. It has been

proposed that the development of RA can be divided into three stages45

. First, autoimmunity

develops based on a complex interplay of genetic and environmental factors. Genetic studies of RA

have identified at least 46 risk loci46

. Especially the human leukocyte antigen (HLA)-DR, protein

tyrosine phosphatase non-receptor type 22 (PTPN22) and peptidyl arginine deiminase type IV

(PADI4) loci, involved in antigen presentation, TCR regulation and citrullination of peptides

respectively, have been associated with RA46-49

. Environmental factors, such as tobacco smoking

and infectious agents have been shown to alter post-translational protein modifications leading to

recognition of autoantigens and the loss of immune tolerance25;50

. This together with the genetic

predispositions can trigger an unwanted immune-response51;52

. Most of the RA patients at this stage

develop autoantibodies, such as rheumatoid factor, anti-citrullinated protein antibody (ACPA) or

anti-collagen type II antibody, which may predict a more aggressive progression of the disease44

. At

the next stage, the inflammation attacks the joint tissue. Activated immune cells accumulate in the

joint and secrete pro-inflammatory cytokines, chemokines and matrix degrading enzymes, including

MMPs, serine proteinases and aggrecanases25;53

. The inflammation increases the formation and

activity of the bone eroding OCs25;43;54

. In the synovial lining resident synovial fibroblasts

proliferate, which causes hyperplasia and pannus formation. The inflamed synovium invades

adjacent cartilage and bone ultimately leading to articular destruction25;44

. Finally, if the local

inflammation is not down-regulated, it develops into a chronic systemic inflammatory condition via

a positive feedback loop that permanently destructs the joints and affects other organs45

.

24

Fig. 3. Schematic representation of a healthy joint (a) and a joint in rheumatoid arthritis (b). In the healthy joint

(a) the thin synovial membrane lines the non-weight-bearing parts of the joint. In rheumatoid arthritis, the synovial

membrane become hyperplastic and immune cells infiltrate the joint. This leads to pannus formation, which invades the

articular cartilage and underlying bone, causing bone erosion and cartilage destruction. Figure from Strand et al. 200755

.

1.2.1 Cellular contribution to rheumatoid arthritis

In RA, cells infiltrating of the synovial tissue, such as monocytes/macrophages, DCs, T- and B cells

play an important role. Monocytes/macrophages are of central importance in RA due to their

prominent numbers in the inflamed synovial membrane and at the cartilage-pannus junction.

Monocytes/macrophages secrete a wide range of proinflammatory cytokines and chemokines,

which recruit and activate other immune cells6. Moreover, they have been reported to release

MMPs, contributing to the local degradation of cartilage25;56

. Monocytes/macrophages are also

thought to be capable of activating T cells in the RA joint thus contributing considerably to the

sustained inflammation and joint destruction in RA57

.

Accumulating evidence show that DCs contribute to the complications of RA58

. Under

non-pathological conditions, DCs are rarely localized adjacent to the bone and they are not likely to

contribute to bone remodeling as it has been shown that DCs-deficient mice have no skeletal

defects59

. However, in the RA synovium, DCs are found to be present in large numbers within the

synovial tissue60

. Various DCs subsets have been identified in the joint of RA patients, contributing

to the induction of T helper 1 (Th1) and Th17 T cell subsets and OCs formation, by expressing

RANK and/or RANKL61;62

. Fc receptors and TLR are highly involved in the modulation of DCs

activity63

. Besides direct contribution of DCs to the osteoclastogenesis by transdifferentiating into

25

OCs (see 1.1.2), they also play an indirect role in bone degradation by producing cytokines such as

IL-1, IL-6 and TNF-, which increase expression and release of cathepsin K and TRAP by resident

OCs8. It has been shown that ColII-pulsed DCs can induce arthritis in a mouse model after adoptive

transfer, indicating a potential role of DCs in driving RA immunopathology64

.

Traditionally, RA has been considered a Th1 cell mediated disorder, driven by the

cytokines IFN- and TNF-25

. Recently, a new model has been proposed, implicating Th17 cells

and their effector cytokines, IL-17, IL-21 and IL-22 as crucial factors for the RA development53;65

.

Th17 cells support osteoclastogenesis, likely through an IL-17-mediated induction of RANKL on

osteoblastic cells and synovial fibroblasts66

. IL-17 facilitates local inflammation by activating

macrophages, monocytes, fibroblasts and endothelial cells leading to release of TNF- and IL-1,

which further enhance the RANKL expression25

. Th1 and Th2 cells have been proposed to inhibit

osteoclastogenesis by acting on the OCs precursor cells, mainly through IFN- and IL-4,

respectively25

.

Fig. 4. IL-17-mediated regulation of RANKL-induced osteoclastogenesis in the RA synovial tissue. Th17 cells

function as a pro-osteoclastogenic Th cell subset by stimulating local inflammation, inducing RANKL on the cells

supporting osteoclastogenesis , and expressing RANKL themselves. It is notable that RANKL on Th17 cells alone is

not sufficient for the induction of OCs differentiation (dotted line). Op, osteoclast precursor cell. Figure from Sato et al.

200665

.

B cells may contribute to RA in several ways, for example by acting as APCs, cytokine producers

and by expressing RANKL. In RA, B cells secrete autoantibodies to citrullinated proteins or anti-Ig

(rheumatoid factor) which lead to formation of immune complexes and complement deposition in

the joint57

. In some RA patients, ectopic germinal centers are formed in the joint. The formation of

germinal centers contributes to B cell maturation and generation of high affinity Abs25;57

. The

importance of B cells in RA has been emphasized by the fact that B cell depleting mAb Rituximab

has shown efficacy in RA patients67

.

26

1.2.2 Cytokines and chemokines in rheumatoid arthritis

Cytokines are directly implicated in the immune processes associated with the pathogenesis of RA.

The cytokines are crucial for maintenance of chronic inflammation and driving the destruction of

joint tissue in the RA patients25

. Elevated levels of proinflammatory cytokines, such as TNF-, IL-

1, IL-6, IL-15, IL-18 and GM-CSF are found in the RA synovium44

. TNF- is a pleiotropic

cytokine that induces the expression of inflammatory factors like cytokines and prostaglandins as

well as enhances OCs differentiation6;43

. TNF- may also induce activation and survival of

leukocytes, endothelial cells and fibroblasts25

. The IL-1 levels in the synovial fluid have been

shown to correlate with the score of the joint inflammation68

. This cytokine appears to mediate the

articular damage to a high degree69

. IL-1 induces MMP-1 and MMP-3 production by macrophages

and fibroblasts, which enhance collagen degradation6;70

. Both TNF-, IL-1 are shown to up-

regulate RANKL on the bone stromal cells, which enhances OCs generation and bone resorption43

.

Overproduction of IL-6 is closely related to the RA pathology71;72

. The level of IL-6 has been found

to be significantly elevated in the synovial fluid from RA patients, especially during the acute

phase73

. IL-6 is secreted by several cell types in the synovium, especially by monocytes/

macrophages, and it can stimulate hepatocytes to produce acute-phase factors. In synergy with

TNF- or IL-1, IL-6 induces the production of vascular endothelial growth factor (VEGF), a

factor important for angiogenesis74

.

Various regulatory cytokines, such as TGF-, IL-10, IL-11 and IL-1RA, are expressed

in the joint. However, they seem not to be present in sufficient amounts to counteract the activities

of the pro-inflammatory cytokines75

. Similarly, the T cell derived cytokines IL-2 and IL-4 are

absent, which may impair generation of regulatory T cells (Treg), and hence would lead to

preferential Th1 and Th17 differentiation6;75

.

Several chemokines have been shown to be elevated in the RA synovial fluid

contributing to the RA pathogenesis. Macrophages secrete IL-8 and MCP-1, which mediate

leukocyte chemotaxis76-78

. T cells and fibroblasts secrete Regulated on Activation Normal T-cell

Expressed and Secreted (RANTES) that attracts T cells to the RA joints76;79

. Fibroblasts,

macrophages and neutrophils secrete Macrophage Inflammatory Protein-1 (MIP-1) that recruits

macrophages and T cells, activates granulocytes and enhances the production of proinflammatory

cytokines80;81

.

27

1.2.3 Rheumatoid arthritis treatment

Different types of treatments exist for patients with RA. The current treatments can improve

symptoms and modify the progression of the disease, but they do not cure RA. The treatments used

for RA include anti-inflammatory agents, analgesics and disease-modifying anti-rheumatic drugs

(DMARDs), such as Methotrexate. DMARDs hinder disease progression, but have a limited effect

and a complete halt of joint destruction is not achieved in the vast majority of the cases82-84

.

Recently, the focus has been on biological agents as alternative treatment options. Especially

targeting cytokines, in particular TNF-, IL-1 and IL-6, has been proposed44

. Blocking TNF- has

been proved successful for treatment of RA and other autoimmune diseases and has highlighted the

importance of inflammation in joint destruction25;40

. Anti-TNF- treatment has been shown to

target inflammation at multiple levels. For instance, it caused decrease in plasma levels of IL-1, IL-

6 and acute-phase proteins, suppressed OCs and DCs maturation as well as leukocyte migration into

the synovium. It has also been shown to led to the recovery of Treg cell formation and function85-88

.

However, a significant number of RA patients show only partial responses or fail to respond to anti-

TNF- treatment25;89

. Furthermore, a whole new type of treatment, consisting of small molecules

that interrupt intracellular signaling through kinase inhibition, is being examined as RA

therapeutics90

. An example of this is Tofacitinib, a JAK inhibitor, which showed promising results

in a phase II study, where it decreased structural damage and improved disease activity in RA

patients91

.

The major challenge in RA management still remains, namely to halt ongoing bone

and cartilage destruction. A promising treatment for halting bone erosion is an anti-RANKL mAb

(Denosumab). Denosumab has been shown to have an anti-resorptive effect in osteoporosis92;93

and

to inhibit structural damage in RA94

. Anti-RANKL therapy might, in combination with anti-

inflammatory drugs, block the disease manifestations.

28

1.3 Extracellular matrix

Extracellular matrix (ECM) is a 3-D structure that surrounds cells and defines their

microenvironment95

. It consists of multiple components of which collagen is the most abundant96

.

There are two main types of ECM, the basement membrane, mainly consisting of collagen type IV

(ColIV), laminins, entactin and proteoglycans, and the interstitial matrix, primarily consisting of

collagen type I (ColI), -type II (ColII), –type III (ColIII) and fibronectins97

. Besides supporting

structural integrity, ECM is important for controlling various cellular functions. Cells use ECM for

homing, migration and invasion. ECM also regulates cell differentiation, function,

activation/maturation, and seems to enhance complement receptor- and Fc receptor mediated

phagocytosis, as well as to stimulate cytokine production98-100

.

1.3.1 Extracellular matrix of bone and cartilage

The human skeleton consists of specialized tissues, such as bone and cartilage. Bone is a dynamic

organ that is continuously being formed, shaped and repaired, involving break down (resorption)

and build-up (synthesis), thus providing maximal strength with minimal mass26

. Bone is composed

of cells and ECM, the latter being further subdivided into an inorganic and organic part. The

inorganic part primarily contains crystals of calcium and phosphorus, whereas the organic part

mainly consists of ColI (approximately 95%), as well as other collagens, non-collagenous proteins

and proteoglycans2. Bone homeostasis depends on the balance between the bone-forming OBs and

the bone-resorbing OCs. The balance is tightly regulated by a number of osteogenic cytokines,

growth factors and hormones101

. Most adult skeletal diseases are due to excess osteoclastic activity,

leading to imbalance in bone remodeling favoring resorption. These include diseases such as

osteoporosis, periodontal disease, RA, multiple myeloma and metastatic cancer26

.

In adults, articular cartilage is an avascular tissue composed of collagens,

proteoglycans and non-collagenous proteins (Fig. 5). The collagen network in cartilage consists of

approximately 60% ColII102;103

. The most prominent proteoglycan in cartilage is aggrecan that

attracts water and gives cartilage protective resiliency. The cellular compartment of cartilage is

constituted of chondrocytes, which regulate cartilage homeostasis104

. The chondrocytes synthesize

new ECM and are also the source of proteinases for the ECM degradation. The turnover of cartilage

is limited and the repair capacity is restricted upon damage104

. Chondrocytes’ function is affected

by the inflammatory environment in RA, where the articular cartilage is damaged mainly by

29

cytokine-induced proteolytic enzymes. Cartilage degradation can be examined by assessing

degradation fragments, such as telopeptides of ColII97

.

Fig. 5. Cartilage components. The articular cartilage is mainly composed of collagen type II, chondrocytes and

aggrecan attached to hyaluronic acid102;103

. Figure borrowed from Suzi H. Madsen.

Collagen

Collagen is the key component of the ECM. It provides essential structural support for connective

tissue and is involved in a spectrum of cellular functions98;99

. All collagens consist of three

polypeptide chains termed chains that are characterized by repeating glycine-X-X’ sequences.

Amino acids in position X and X’ are often proline and 4-hydroxyproline (O), respectively105

. The

three chains are twisted around each other into a right-handed super helix. In vertebrates, 28

different types of collagen have been identified, composed of at least 46 distinct collagen

polypeptide chains106

. Several collagens carry glycosaminoglycans chains and are also considered

proteoglycans107

. Collagens can assemble into supramolecular structures, such as fibrils, which are

stabilized by the formation of covalent cross-links. ColI and ColII belong to the group of classical

fibril forming collagens107

. Collagen interacts with cell surface receptors. Some binding sites are

only available for interaction after denaturation and these motifs may be important in the removal of

degraded and denatured collagen or in autoimmune responses99

.

30

Fig. 6. Schematic representation of the structure of collagen. (a) Surface representation of a collagen triple helix, as

a model for the collagen-like peptide (PPG)9. The three α-chains wind around one another with a one-residue stagger.

(b) Schematic representation of how individual collagen triple helices assemble into a quarter-staggered array to form a

collagen fibril. (c) Electron micrograph of a heterotypic collagen fibril isolated from human articular cartilage. Figure

from Leitinger, B. 201199

.

1.3.2 Extracellular matrix turnover

The pleiotropic ECM functions are defined by a highly dynamic structure of ECM and its

remodeling capabilities to control cellular behaviour108

. Four classes of enzymes can degrade matrix

proteins: cysteine, aspartate- and serine-dependent proteases, and metalloproteinases. The cysteine

and aspartate-dependent proteases act primarily intracellularly, while serine-dependent proteases

and metalloproteinases act extracellularly109

. Connective tissue destruction is majorly mediated by

metalloproteinases. Within the metalloproteinases classes, two main families exist: the matrix

metalloproteinase (MMP) and a disintegrin and metalloproteinase with trombospondin motifs

(ADAMTS)108

. Native fibrillar collagens are resistant to proteases such as pepsin, trypsin or

chrymotrypsin98

.

Matrix metalloproteinases

MMPs are zinc-dependent endopeptidases which currently comprise 23 related, but distinct

proteins110;111

. MMPs are classified into five subtypes: 1) the gelatinases, which degrade ColIV and

other basement membrane proteins; 2) the collagenases, which degrade the fibrillar collagens (ColI,

ColII and ColIII); 3) the stromelysins, which degrade non-collagen matrix-proteins; 4) membrane-

type MMPs and 5) a diverse subgroup110;112

. Fibril-forming collagens (I-III) are cleaved by MMP-1,

-8 and 13. Furthermore, MMP-2 and 14 are capable of cleaving ColI107

.

31

In steady-state tissue MMP activity is very low. However, the expression of the

MMPs can be regulated at the transcription level by several factors, such as cytokines, hormones,

cell-cell interactions and cell interactions with ECM113;114

. The MMPs are secreted as inactive pro-

MMPs that need to be activated. They are converted into the active form by cleavage of their pro-

domain. The mode of MMP activation is not fully elucidated, but they can be activated by other

active MMPs, oxidants and by plasmin110;115;116

. The active MMPs can be inhibited by

internalization or by inhibitors, like tissue inhibitors of metalloproteinases (TIMPs)117

.

Plasmin

Additional mediators of ECM degradation are the serine proteases, like plasmin, elastase and

cathepsin G109

. The zymogen plasminogen (plg) is mainly produced by the liver and circulates in

the plasma at high concentrations (2μM). Plg is converted to its active form, plasmin, by tissue Plg

activator (tPA) or urokinase Plg activator (uPA). When converted, plasmin binds to the ECM and

degrades multiple ECM proteins118;119

. Plasmin can also activate other ECM-degrading proteinases,

such as MMP-3, MMP-9, MMP-12, and MMP-13120

. Plg-dependent activation of MMP-9 has been

shown to be required for inflammatory macrophage migration across ECM121

. The Plg system is

regulated at several levels and potent inhibitors are Plg activator inhibitor–1 (PAI-1), PAI-2, α2-

antiplasmin and aprotinin118;119

. Urokinase-type Plasminogen Activator Receptor Associated Protein

(uPARAP) has been shown to be important for degradation of collagen by macrophages122

.

32

Fig. 7. Plasminogen activating system. The uPA receptor (uPAR) binds both pro-uPA and active uPA, thus restricting

the activation of plasminogen to the cell membrane. When pro-uPA is bound to the uPAR it facilitates plasmin-

mediated activation of uPA. Activated uPA then cleaves plasminogen; generating active plasmin, creating a positive

feed-back loop. Plasmin cleaves and activates MMPs and together they can degrade ECM. Plasminogen can also be

converted to its active form by tissue plasminogen activator (tPA). The major inhibitors of the Plg system are Plg

activator inhibitor–1 (PAI-1), PAI-2, aprotinin and α2-antiplasmin. Figure from Rao, J. 2003121;123

.

1.3.3 Collagen receptors

A structurally and functionally diverse group of surface receptors mediates recognition of collagen,

such as integrins, discoidin domain receptors (DDRs), glycoprotein VI (GPVI), and leukocyte-

associated immunoglobulin-like receptor-1 (LAIR-1) (Fig. 8). These receptors regulate cell

adhesion and migration, homeostasis and immune functions99

. The importance of proper

collagen:collagen-receptor interaction is emphasized by their involvement in pathogenesis of

various diseases, where dysregulation of expression of collagens or collagen-receptors contribute to

disease pathology, such as fibrosis, wound healing, inflammatory disorders and tumor

angiogenesis99

.

Integrins are the main family of collagen receptors and primarily mediate cell

adhesion124

. Integrins are heterodimers composed of non-covalently associated and subunits,

forming 24 distinct integrins. The α1β1 and α2β1 integrins are the most broadly expressed collagen-

binding integrins. The discoidin domain receptors, DDR1 and DDR2, constitute a subfamily of

receptor tyrosine kinases that are characterized by the presence of a discoidin domain in their

33

extracellular region. DDRs form constitutive dimers in the cell membrane99

. DDRs regulate

important cellular processes such as proliferation, migration and differentiation125

. In smooth

muscle cells and fibroblasts DDRs are found to influence remodeling of ECM through the induction

of MMPs126;127

. Both integrins and DDRs display exquisite specificity in their interaction with

collagen by binding specific amino acid side chains in the collagen triple-helix. Their binding

motifs are distinct and non-overlapping99

. The structurally related receptors GPVI and LAIR-1 have

similar collagen-binding mode but mediate opposing functions128

. GPVI is an activating receptor on

platelets and LAIR-1 is an inhibitory receptor on immune cells99

. GPVI is central to haemostasis

where it activates platelets upon binding to collagen129

. LAIR-1 contributes to the regulation of the

immune system by delivering inhibitory signals via its ITIM signaling motif130

. Besides LAIR-1,

LAIR-2 is a part of the subfamily and is a soluble homolog of LAIR-1. It is hypothesized that

LAIR-2 may function as an inhibitor of LAIR-1 signaling in vivo99

.

Recently, a potential crosstalk between different collagen receptors have been

studied99

. Xu and colleagues have reported that the DDRs positively modulate integrin-mediated

cell adhesion131

. Mocsai and co-authors found that activation of integrins was decreased when

ITAM signaling was defective31

. Moreover, ITIM-associated receptors are potent inhibitors of

ITAM-signaling. It has been observed that LAIR-1 inhibits signals by ITAM-bearing receptors

leading to suppression of T cell activity, down-regulation of B cell function and blocking of

cytokine-induced signals99

.

Fig. 8. Schematic domain structures of the collagen receptors. Four classes of collagen binding transmembrane

receptors, the collagen-binding integrins, discoidin domain receptors (DDRs), glycoprotein VI (GPVI) and leukocyte-

associated immunoglobulin-like receptor-1 (LAIR-1) Figure from Leitinger, B. 201199

.

34

Extracellular matrix receptors on myeloid cells

The cells of the myeloid lineage have a large number of receptors regulating their activity. Several

studies have addressed the involvement of the ECM proteins in the control of the immune cell

functions. It has been observed that collagens stimulate maturation of DCs10

as well as induce

cytokine production by PBMCs100

and RA synovial fluid cells132

.

In recent years, a growing number of receptors of the lectin and immunoglobulin

superfamilies (IgSFs), linked to ITAMs/ITIMs signaling, have been described133

. ITAMs are mainly

found in the cytoplasmic tail of transmembrane adaptors, such as CD3, FcR or DAP12, and play a

role of signaling subunits when associated with cell surface receptors. ITAM signaling induces

cellular activation via a phosphorylation cascade triggered by protein tyrosine kinases134

. The

transmembrane adaptors DAP12 and FcR, associated with TREM2 and OSCAR receptors,

respectively, have both been identified to be involved in OCs and DCs differentiation and

function29;32;135;136

. Combined with the increasing evidence that differentiation of OCs occurs

locally in the inflamed synovial membrane, where OCs can form from monocytic precursor cells, a

growing number of studies was focused on the monocyte/OCs receptors TREM2 and OSCAR. My

project is focused on the biology of human OSCAR in various immune cells present in the RA

joint.

35

1.4 Osteoclast-associated receptor (OSCAR)

1.4.1 Murine OSCAR

Osteoclast associated receptor (OSCAR), was initially identified in mice by Kim and colleagues,

and characterized as a novel member of the immunoglobulin (Ig)-like surface receptor family29

. In

mice, OSCAR has been found to be specifically expressed on osteoclast precursors and mature

osteoclasts29

. OSCAR is a monomeric N-glycosylated cell-surface protein of approximately 45 kDa.

The protein exhibits two C2-type immunoglobulin extracellular domains, a transmembrane domain

that contains a positively charged arginine residue, and a short intracellular domain. OSCAR is

associated with the ITAM harboring Fc receptor -chain (FcR) that serves as a signaling subunit of

the receptor complex135

. OSCAR has been recognized as a co-stimulatory receptor, contributing to

osteoclastogenesis by activating NFATc1, an important transcription factor of OCs

differentiation29;30;135

.

The role of OSCAR and its ITAM-harboring adaptor have been examined in murine

models. FcR-/-

mice show no significant difference compared to wild-type mice in respect to bone

disorders. Additionally, the OCs in the knock-out mice were well developed in the presence of M-

CSF and RANKL, and when cultured in the presence of OBs, suggesting that FcR is not essential

for the differentiation of OCs135

. However, several groups have reported that mice deficient in both

DAP12 and FcR exhibit severe osteopetrosis and show defective differentiation of osteoclasts

31;32.

This cannot simply be attributed to DAP12 deficiency, since single knock out DAP12-/-

mice only

exhibit mild osteopetrosis and normal numbers of OCs in bone tissue. Co-culturing DAP12-/-

OC

precursors with OBs eradicated the effect of knocking out DAP1231

. These results suggest that

OSCAR may co-stimulate osteoclastogenesis independent of TREM2-DAP12 signaling137

and that

DAP12 and FcR functionally compensate for the lack of each other32

.

36

Fig. 9. A schematic model of ITAM-mediated co-stimulatory signal in RANK-stimulated induction of OC

differentiation. ITAM-stimulation by immunoreceptors and RANKL–RANK interaction results in NFATc1 induction.

RANKL may also contribute to efficient ITAM signaling through the induction of immunoreceptors. Figure from Koga

et al.32

1.4.2 Human OSCAR

Human OSCAR (hOSCAR) was identified by Merck et al.136

. The murine and human OSCAR

share approximately 73% amino acid identity and the genes can be regarded as orthologs29

. In

humans, OSCAR is more widely expressed than murine OSCAR and is present on all cells of the

myeloid cell lineage136

. Like in mice, hOSCAR is associated with FcR and can signal to activate

myeloid cells by triggering calcium influx and cytokine release136

. The calcium influx induces the

master transcription factor for osteoclastogenesis, NFATc132

, shown in Fig. 9. The gene coding for

OSCAR, known as PIGR3, has been mapped to chromosome 19q13136

. OSCAR exists as both

membrane-bound and soluble isoforms138-140

presumably resulting from alternative splicing event29

and/or being proteolytically shed from the cell membrane141

.

1.4.3 OSCAR function

OSCAR has diverse functions, playing a role in both bone homeostasis and activation of the

immune response. OSCAR serves as a co-stimulatory molecule for OCs differentiation, acting in an

additive manner with primary stimulation for RANKL, enabling a fine-tuning of OCs

differentiation29;137;141

. Moreover, OSCAR has been suggested to be implicated in the development

and maturation of cells of the myeloid lineage11;134;136

. Most studies on the hOSCAR function in

myeloid cells have been performed using stimulation of the OSCAR signaling by crosslinking of

OSCAR with specific antibodies, since a naturally occurring ligand of OSCAR was not identified

until recently137;142

. Merck and colleagues reported that crosslinking OSCAR on iDCs led to

activation and maturation. The cells were up-regulating the phenotypic markers CD40, CD54,

37

CD80, CD86, HLA-DR and partially CD8311

. Up-regulation of activation and maturation markers

was also observed for monocytes, and to a lesser extent on neutrophils134

.

Crosslinking of hOSCAR in iDCs has been shown to induce high expression levels of

the chemokines TARC/CCL17 and MDC/CCL22. These chemokines are known to be able to attract

Th2 effectors and regulatory T cells143;144

. Moreover, cross-linking of OSCAR on DCs, monocytes

and neutrophils induced secretion of IL-8 and MCP-1, which are involved in the recruitment of

leukocytes to sites of inflammation11

. Furthermore, ligation of OSCAR on neutrophils caused

degranulation, respiratory burst activity and release of antibacterial molecules, namely

myeloperoxidase, MMP-9 and lactoferrin134

.

OSCAR stimulation has further been suggested to promote survival of monocyte-DCs

in the absence of survival factors by maintaining expression of anti-apoptotic molecules of the Bcl-

2 family through activation of the PI3K and ERK signaling pathways11

. The same was observed for

monocytes, but not for neutrophils134

. Merck and colleagues have suggested a role of OSCAR in

antigen endocytosis and presentation in DCs. They demonstrated that upon cross-linking of OSCAR

with an anti-OSCAR mAb the whole receptor/mAb complex was endocytosed. The internalized

mAb peptides reached vesicles characteristic of the lysosomal/late endosomal MHC II-loading

compartment. The peptides derived from the anti-OSCAR mIgG1 mAb were presented to T cells

and stimulated proliferation of the mIgG1 specific T cell clone136

. Merck and colleagues also

observed that OSCAR signaling in DCs could modulate the response of the cells to the Toll-like

receptors (TLR) engagement11

. Co-stimulation of DCs with the OSCAR surrogate ligand

(crosslinking mAb) and TLR ligands, but not by OSCAR cross-linking alone, synergistically

enhanced pro-inflammatory cytokine release and maturation of DCs as well as the ability of the

DCs to induce proliferation of naïve T cells 11

. Synergistic action of OSCAR and TLR ligands was

also observed in monocytes and neutrophils134

.

1.4.4 OSCAR ligand

Kim and colleagues found that an OSCAR-IgFc fusion protein bound to the surface of osteoblasts

indicating that these cells express the ligand of OSCAR29

. The group of Trowsdale then reasoned,

based on the structural homology of OSCAR and collagen receptors GPVI and LAIR1, that the

OSCAR ligands could be ECM proteins and hence screened a library of the triple-helical ColII

peptides (ToolKit II). OSCAR was then identified as a receptor for ColII and was further shown to

bind strongly to collagen I, II and III, weakly to collagen IV, and not to collagen V137

. OSCAR did

not bind to the triple-helical peptide ligands for integrin 21, the GPVI ligand, (GPP)10 control

38

peptides or to the ECM proteins vitronectin or fibronectin137

. Specific binding of collagen to

OSCAR was shown by adding an anti-hOSCAR-blocking mAb that inhibited the interaction137

.

Functionally, the authors were capable of inhibiting the collagen peptide-induced osteoclastogenesis

by adding anti-hOSCAR blocking mAb to the monocytes cultured in the presence of RANKL,

supporting that the co-stimulatory signaling effect of collagen peptides are hOSCAR specific137

.

1.4.5 Regulation of OSCAR

Regulation of OSCAR has not been fully elucidated. The OSCAR gene expression was shown to be

induced by NFATc1 as well as by the microphtalmia-inducing transcription factor (Mitf) that works

synergistically with the transcription factor PU.1145

. The expression of murine OSCAR was shown

to be induced during RANKL-induced osteoclastogenesis29

through stimulation of NFATc1, which

initiates transcription of OCs-specific genes, indicating a positive feedback loop between OSCAR

and NFATc130

. It has also been reported that OSCAR expression is elevated on monocytes from

RA patients stimulated with TNF- but not with RANKL or M-CSF141

.

OSCAR signaling is negatively regulated by ITIM receptors such as CD85j (leukocyte

Ig-like receptor-1/Ig-like transcript 2)12

and LAIR-1 (Leukocyte-Associated Immunoglobulin-Like

Receptor 1)99

. For example, Tenca and colleagues12

showed that numerous DCs functions triggered

by hOSCAR stimulation were negatively regulated by CD85j. In DCs, CD85j has been shown to

inhibit intracellular calcium mobilization and strongly reduce OSCAR-mediated secretion of IL-8

and IL-12(p40). Moreover, CD85j was able to counteract the anti-apoptotic effect of OSCAR, by

reducing the Bcl-2 expression enhanced by OSCAR stimulation12

. These observations are in line

with the findings that osteoclastogenesis is enhanced in mice lacking SHP-1 or SHIP-1

phosphatases, which counterbalance the ITAM-signal in the immune system32

.

1.4.6 OSCAR in disease

OSCAR has been linked to bone disorders as well as inflammatory diseases. A single nucleotide

polymorphism within the OSCAR gene has been associated with low bone mass and osteoporosis in

postmenopausal women146

. Herman and colleagues were the first to address the role of OSCAR in

RA, looking into the OSCAR expression in the synovial tissue of RA patients. OSCAR was

expressed by OCs at the bone erosion front and by mononuclear cells (MNCs) surrounding synovial

micro vessels141

. The OSCAR expression was also associated with microvasculature in patients

with RA and OA, which is not seen in healthy tissue139

. Additionally, it was observed that the

OSCAR expression was enhanced in peripheral blood monocytes of RA patients as compared to

39

healthy subjects141

. The expression level of OSCAR positively correlated with disease activity as

well as with the potency of monocytes to differentiate into OCs in vitro, specifying the importance

of OSCAR as an enhancer of osteoclastogenesis.

Contradictory results have been shown for soluble OSCAR (sOSCAR) in synovial

fluid and plasma from RA patients138-140

. One study by Zhao et al. showed that serum levels of

sOSCAR were significantly lower in RA patients than in healthy controls138

. Opposite data was

seen in a study including 136 RA patients, who had significantly higher plasma levels of sOSCAR

compared to the healthy controls. The level of sOSCAR was also shown to correlate with the

disease progression, as the patients with destructive RA had significantly higher amounts of

sOSCAR compared to non-destructive RA patients140

.

OSCAR was also reported to be expressed on endothelial cells147;148

. In endothelial

cells, OSCAR expression is increased by oxidized low-density lipoproteins147

. Further analysis of

the OSCAR signaling in the endothelial cells revealed its involvement in proliferation,

inflammatory response and cell-cell signaling148

. Moreover, OSCAR expression in monocytic cells

was shown to be induced in atherosclerotic mice and was suggested to be regulated by pro-

atherogenic stimuli, indicating that OSCAR may play a role in the pathogenesis of

atherosclerosis149

.

40

2. Rationale and aim

As a part of the NN project “OSCAR intervention for treatment of RA”, the PhD project aimed to

gain further insights into the pathogenesis of RA. One of the characteristic features of rheumatoid

arthritis is the presence of immune cells in the synovium. Immune cells, such as DCs, monocytes

and macrophages are shown to express OSCAR136

. OSCAR has been shown to serve as a co-

stimulatory molecule for OCs differentiation29;137

and a functional role of OSCAR in human

myeloid cells have been proposed by Merck and colleagues11;134;136

. Collagen was recently

identified as naturally occurring OSCAR ligand137;142

and the functional implication of OSCAR-

collagen interaction in myeloid cells was not addressed so far. In the current project, we aimed to

assess the role of OSCAR-collagen signaling in DCs, monocytes and macrophages, using an in

vitro approach with human primary cells. Since collagen is the major component of the bone and

cartilage, it was hypothesize that OSCAR engagement by collagen on myeloid cells could

contribute to the RA pathogenesis as a driver of sustained inflammation, cartilage damage and

inflammatory bone resorption. The project therefore also intended to evaluate the effect of blocking

OSCAR-collagen interaction in DCs, monocytes and macrophages.

The outcome of the work was anticipated to provide new knowledge of the mechanisms controlling

synovial inflammation, inflammatory bone erosion and cartilage degradation, with the perspective

of identifying new therapeutic targets.

41

3. Results

The aim for this PhD thesis is to examine the functional role of the OSCAR-collagen interaction in

human immune cells of the myeloid origin.

The results are presented in the following three parts:

Results part 3.1: exploring the role of OSCAR in DCs

Results part 3.2: studies on the OSCAR-collagen interaction in monocytes from healthy donors

and synovial fluid cells from RA patients

Results part 3.3: combines preliminary data on the role of OSCAR in macrophages

42

3.1 OSCAR function in dendritic cells

This part of my studies has been published in:

Heidi S. Schultz, Louise M. Nitze, Louise H. Zeuthen, Pernille Keller, Albrecht Gruhler, Jesper

Pass, Jianhe Chen, Li Guo, Andrew J. Fleetwood, John A. Hamilton, Martin W. Berchtold and

Svetlana Panina. (2015) Collagen Induces Maturation of Human Monocyte-Derived Dendritic

Cells by Signaling through Osteoclast-Associated. J. Immunology. 194 (epub, doi 10.4049/

jimmunol.1402800)

For supplemental data see Appendix I

The Journal of Immunology

Collagen Induces Maturation of Human Monocyte-DerivedDendritic Cells by Signaling through Osteoclast-AssociatedReceptor

Heidi S. Schultz,*,† Louise M. Nitze,*,† Louise H. Zeuthen,* Pernille Keller,*

Albrecht Gruhler,* Jesper Pass,* Jianhe Chen,‡ Li Guo,‡ Andrew J. Fleetwood,x

John A. Hamilton,x Martin W. Berchtold,† and Svetlana Panina*

Osteoclast-associated receptor (OSCAR) is widely expressed on human myeloid cells. Collagen types (Col)I, II, and III have been

described as OSCAR ligands, and ColII peptides can induce costimulatory signaling in receptor activator for NF-kB–dependent

osteoclastogenesis. In this study, we isolated collagen as an OSCAR-interacting protein from the membranes of murine osteo-

blasts. We have investigated a functional outcome of the OSCAR–collagen interaction in human monocyte-derived dendritic cells

(DCs). OSCAR engagement by ColI/II-induced activation/maturation of DCs is characterized by upregulation of cell surface

markers and secretion of cytokines. These collagen-matured DCs (Col-DCs) were efficient drivers of allogeneic and autologous

naive T cell proliferation. The T cells expanded by Col-DCs secreted cytokines with no clear T cell polarization pattern. Global

RNA profiling revealed that multiple proinflammatory mediators, including cytokines and cytokine receptors, components of the

stable immune synapse (namely CD40, CD86, CD80, and ICAM-1), as well as components of TNF and TLR signaling, are

transcriptional targets of OSCAR in DCs. Our findings indicate the existence of a novel pathway by which extracellular matrix

proteins locally drive maturation of DCs during inflammatory conditions, for example, within synovial tissue of rheumatoid

arthritis patients, where collagens become exposed during tissue remodeling and are thus accessible for interaction with infil-

trating precursors of DCs. The Journal of Immunology, 2015, 194: 000–000.

Rheumatoid arthritis (RA) is a systemic autoimmune dis-ease that primarily targets the peripheral diarthrodialjoints. It is characterized by the presence of autoanti-

bodies, synovial inflammation, pannus formation, cartilage dam-age, and bone erosion. The molecular mechanisms involved in RApathogenesis remain obscure. A multistep mechanism of RA de-velopment has been proposed (1, 2). First, unknown environmentalfactors cause alterations in posttranslational protein modifications,leading to recognition of autoantigens and the loss of immune

tolerance. Next, a tissue-specific inflammatory response is elicitedin the joints. Activated immune cells accumulate in the synovialtissue and secrete cytokines, chemokines, and matrix-degradingenzymes, causing tissue damage and remodeling. By way ofa positive feedback loop, the local inflammation is sustained andfinally progresses into a systemic disorder.The cellular infiltrates in the synovial tissue are mainly com-

posed of blood-derived cells, including T and B cells, as wellas macrophages and dendritic cells (DCs) differentiated frommonocytes in the RA synovial microenvironment. Cells of themyeloid lineage are regarded as pivotal regulators of RA. Bothimmature and mature DC subsets are present in large numberswithin the RA joint, and strong evidence supports the importance ofDCs in synovial inflammation (3, 4). DCs are professional APCsthat are highly capable of inducing T cell responses. In RA, theDCs are essential for initiation and perpetuation by Ag presenta-tion as well as for cytokine and chemokine secretion (5, 6). Fur-thermore, DCs are capable of transdifferentiating into osteoclasts,contributing directly to bone erosion (7, 8). The maturation stateof the DCs is crucial for determining the balance between toler-ance and immunity. Under homeostatic conditions, DCs arelargely immature and are thought to induce tolerance. Changes inthe microenvironment stimulate maturation of DCs, which canpromote immunogenicity.A human ortholog of murine OSCAR was identified by Merck

et al. (9). OSCAR is expressed on cells of myeloid origin, and itsignals through the ITAM-harboring adaptor protein FcRg (9, 10).In DCs, cross-linking of OSCAR with specific mAbs enhancedAg presentation and promoted a semimature phenotype, causingchemokine but not proinflammatory cytokine secretion. In synergywith TLR signaling, but not alone, OSCAR cross-linking en-hanced the ability of DCs to induce naive CD4+ T cell prolifer-

*Biopharmaceutical Research Unit, Novo Nordisk A/S, 2760 Maløv, Denmark;†Department of Biology, Copenhagen University, Copenhagen 2200, Denmark;‡Novo Nordisk Research Center China, Beijing 102206, China; and xDepartmentof Medicine, University of Melbourne, Royal Melbourne Hospital, Parkville, Victoria3050, Australia

Received for publication November 5, 2014. Accepted for publication January 22,2015.

H.S.S. and L.M.N. received Ph.D. fellowships from the Danish Agency for Science,Technology, Innovation and Research.

The microarray dataset presented in this article has been submitted to the ArrayExpress database (http://www.ebi.ac.uk/arrayexpress/experiments/E-MTAB-2904/)under accession number E-MTAB-2904.

Address correspondence and reprint requests to Dr. Svetlana Panina at the currentaddress: Novo Nordisk A/S, Hvidkildevej 111, 2400 Copenhagen, Denmark. E-mailaddress: [email protected]

The online version of this article contains supplemental material.

Abbreviations used in this article: Col, collagen type; Col-DC, collagen-matured DC;DC, dendritic cell; ECD, extracellular domain; ECM, extracellular matrix; FDR,false discovery rate; iDC, immature DC; LC-MS, liquid chromatography–mass spec-trometry; MS, mass spectrometry; NP-40, Nonidet P-40; OSCAR, osteoclast-associated receptor; RA, rheumatoid arthritis; SPR, surface plasmon resonance.

This article is distributed under The American Association of Immunologists, Inc.,Reuse Terms and Conditions for Author Choice articles.

Copyright� 2015 by The American Association of Immunologists, Inc. 0022-1767/15/$25.00

www.jimmunol.org/cgi/doi/10.4049/jimmunol.1402800

ation (9, 11, 12). It was concluded that OSCAR can inducea semimatured phenotype in DCs, which is characterized by ex-pression of costimulatory molecules but an inability to secreteproinflammatory cytokines. Such semimatured DCs are implicatedin the induction of tolerance.Recent studies have assessed the role of OSCAR in RA. OSCAR

is expressed on mononuclear cells surrounding synovial micro-vessels and on multinucleated giant cells at the bone resorptionareas. OSCARwas also found to be upregulated in peripheral bloodmonocytes from RA patients compared with healthy subjects, andits expression correlated to disease activity (13).In the present study we aimed to examine OSCAR function in

DCs with relevance to RA pathogenesis. We identified collagentypes (Col)I/II as OSCAR ligands and investigated the func-tional outcomes of OSCAR–collagen interaction in DCs. Duringthe preparation of this manuscript, Trowsdale and colleagues (14)identified ColI, -II and -III to be ligands of OSCAR (14). Theauthors showed that within joints collagens are exposed toOSCAR-expressing osteoclasts and osteoclast precursors. In vitro,OSCAR-binding collagen peptides promoted receptor activator forNF-kB ligand–induced osteoclast differentiation from blood-derived monocytes, indicating that OSCAR is an activating re-ceptor on osteoclast precursor cells. In the present study, we showthat OSCAR–collagen interaction provides an activating signalalso in monocyte-derived DCs. This indicates an important, yetpoorly recognized, role of the extracellular matrix (ECM) in thearthritic joint, where ColI/II become exposed due to ongoing tis-sue remodeling, thus providing activating signals to myeloid cellsto drive a sustained inflammation. Blocking the OSCAR signaling,for example, with a mAb, may be considered as a novel thera-peutic approach for RA.

Materials and MethodsProteins and Abs for functional assays

The extracellular domain (ECD) of human OSCAR and OSCAR–Fc fusionprotein were produced by recombinant technique as follows. The fragmentencoding ECD (amino acid residues 1–227; sequence ID NP_573399) wasamplified by RT-PCR using human universal cDNA (Clontech) as tem-plate, cloned into the pJSV002 expression vector, and sequenced. Therecombinant protein was expressed in HEK293-6E cells. The purifiedrecombinant protein is an N-glycosylated monomer of ∼25 kDa. For ex-pression of OSCAR–Fc protein, the ECD-encoding cDNA was cloned in-frame with a sequence encoding the Fc fragment of human IgG4. Thesequence was verified and the protein was expressed in HEK293-6E cells.The purified protein is an N-glycosylated dimer of ∼100 kDa. For cellbinding studies, OSCAR-ECD was labeled using an Alexa Fluor 488protein staining kit (Life Technologies) according to the manufacturer’sprotocol.

A panel of murine mAbs against OSCAR-ECD and isotype controlswere generated using hybridoma technology and then humanized usingrecombinant technology. The CDR of the generated mAb was grafted ona germline-encoded, human-mutated IgG1 Fc part, Fc-hIgG1.1 (15, 16).Five point mutations introduced in the Fc-hIgG1.1 hinder its binding tohigh-affinity Fc receptors. As isotype controls, an anti-OSCAR non-blocking high-affinity mAb as well as anti-TNP hIgG1.1 were used. Mostblocking experiments were performed with the use of several differenthigh-affinity anti-OSCAR–blocking mAbs and the two above-mentionedisotype controls with identical results. For simplicity, data are shownonly for 1 mg/ml mAb unless specified otherwise. For functional assays,the following commercially available mAbs were used: anti-CD3 mAb(clone OKT3; eBioscience), anti-CD3/CD28 expander beads (Dynal),and anti–integrin a2b1 mAb (clone BHA2.1; Millipore). ColI purifiedfrom human skin was purchased from Sigma-Aldrich. ColII purifiedfrom human cartilage was purchased from Millipore. ColIII purified fromhuman placenta was purchased from Millipore. ColIV from humanplacenta was purchased from Sigma-Aldrich. Endotoxin level in theprotein solutions was measured using the turbidimetric kinetic Limulusamebocyte lysate assay (reagents supplied by Charles River Laborato-ries). All proteins used for functional assays in this study were essen-tially endotoxin-free.

Identification of OSCAR ligand

OSCAR-binding proteins were purified from membranes of MC3T3-E1cells by affinity chromatography on an OSCAR–Fc matrix followed byidentification of hits using a liquid chromatography–mass spectrometry(LC-MS) approach. In brief, for extraction of membrane proteins, the cellswere washed in PBS, scraped with a plastic policeman, and homogenizedmechanically using a glass pastel homogenizer in the buffer containing 10mM imidazole and 1 mM EDTA (pH 7.4) on ice. The homogenate wasspun at 20,000 3 g for 10 min, and the pellet was resuspended in thehomogenization buffer. This washing cycle was repeated three times withone last time in a buffer containing 2 mM imidazole/1 mM EDTA (pH7.4). The pellet was then washed twice in 50 mM Tris-HCl/150 mM NaCl(pH 8.2), centrifuged, and resuspended in a RIPA buffer (150 mM NaCl,50 mM Tris-HCl [pH 8.2], 1 mM EDTA, 1% Nonidet P-40 [NP-40], 1 mg/mlBSA, protease inhibitor [cOmplete, Roche] and phosphatase inhibitormixture [Sigma-Aldrich]) for 30 min on ice. The protein extracts wereclarified by centrifugation at 20,000 3 g. Protein concentration wasmeasured using a bicinchoninic acid protein assay (Pierce).

Preparation of an affinity resin was performed as follows. HumanOSCAR–Fc IgG4 or Fc IgG4 (control) recombinant proteins were bound toa protein G–Sepharose 4B Fast Flow affinity matrix (Sigma-Aldrich)according to the manufacturer’s protocol and then chemically cross-linked as described in Schneider et al. (16). This technique allows opti-mal orientation of OSCAR-Fc on the beads and thus efficient binding ofinteracting proteins with minimal/no leakage of the bait during the elutionprocess. After cross-linking, the beads were thoroughly washed in 0.1 Mborate buffer, equilibrated in 5 ml 0.5 M Tris-HCl (pH 8.2), and sedi-mented by centrifugation. Cell membrane extracts in RIPA buffer wereadded to the affinity beads. NaCl concentration was adjusted to final 0.5 Mand the beads were incubated for 3 h at 4˚C with gentle agitation. Thebeads were then washed sequentially with three buffers containing 1) 0.5 MNaCl, 50 mM Tris-HCl, 1 mM EDTA, and 0.5% NP-40 (pH 8.2); 2) 150mM NaCl, 50 mM Tris-HCl, 0.5% NP-40, and 0.1% SDS (pH 8.2); and 3)150 mM NaCl and 0.5% CHAPS. The latter wash was repeated twice. Theproteins bound to the affinity matrix were eluted in 50 mM diethylamine/0.5% CHAPS (pH 11.5) and pH was immediately adjusted to neutral usingPBS and borate buffer. The samples were subjected to a 4–12% SDS-PAGE. The gels were stained using a ProteoSilver plus silver stain kit(Sigma-Aldrich). The bands differentially bound to the OSCAR–Fc matrixwere subjected to LC-MS analysis as follows. Bands were excised from theSDS-PAGE gel, destained with the SilverQuest silver staining kit (Invi-trogen), reduced with DTT, alkylated with iodoacetamide, and digestedwith trypsin with a slightly modified version of the method described inShevchenko et al. (17).

LC-MS analysis

Peptide mixtures were analyzed by nanoscale liquid chromatography–electrospray ionization–tandem MS on an Easy-nLC (Proxeon Biosystems)connected to an LTQ Orbitrap mass spectrometer (Thermo Fisher Scien-tific). Reversed phase chromatography was performed with 0.1% formicacid (buffer A) and acetonitrile/0.1% formic acid (buffer B). Peptides weretrapped on a precolumn (Proxeon Biosystems) and separated on a 10-cmfused-silica column (internal diameter 75 mm) packed with 3-mmReproSil-Pur C18-AQ (Proxeon Biosystems). MS analysis was performedon the LTQ Orbitrap in data-dependent mode, where one MS survey scanwas followed by MS fragmentation spectra of the four most intense ions.

Analysis of the MS data were performed with Proteome Discoverer 1.4(Thermo Fisher Scientific). Peak lists of the MS fragmentation spectra weregenerated and searched in MASCOT against the mammalian part of theSwissProt database with oxidation (P), oxidation (M), and deamidated (NQ)as dynamic modifications and carbamidomethyl (C) as fixed modification.Results were filtered for a false discovery rate (FDR) of 0.01 and aminimumMASCOT peptide score of 25.

Binding of OSCAR to collagen

Binding of the soluble extracellular domain of OSCAR to human collagenswas tested by surface plasmon resonance (SPR) on a ProteOn instrument(Bio-Rad). Human collagens were immobilized on a GLC sensor chip usingthe amine coupling kit (Bio-Rad). The experiment was performed at 15˚Cusing HBS-EP+ (GE Healthcare) as running buffer for the immobilizationand HBS-EP+/0.1% BSA for binding analysis. The chip was activated with1-ethyl-3-(3-dimethylaminopropyl)carbodiimide hydrochloride/normalhuman serum (diluted 2:1 with H2O) according to instructions. Collagenswere diluted in 10 mM NaAc buffer (pH 4.5) to 25 mg/ml and injected for200 s followed by quenching with ethanolamine. OSCAR-ECD wasinjected at 2 mM, 1 mM, 500 nM, 250 nM, and 125 nM for 200 s at a flow

2 OSCAR-DEPENDENT MATURATION OF DCs

of 60 ml/min followed by regeneration with 10 mM glycine (pH 2.0) for30 s. Binding curves were double referenced by subtraction of the signalsfrom reference cells and a buffer injection using the ProteOn Managersoftware.

Competition of OSCAR binding to collagen

Competition of anti-OSCAR mAb with OSCAR binding to collagen wasmeasured by SPR on a Biacore T100 instrument using the same buffers asdescribed above.

Human ColI was diluted with 10 mM NaAc (pH 4.5) to a final con-centration of 2.5 mg/ml and immobilized on a CM5 sensor chip (GEHealthcare) using the amine coupling kit (GE Healthcare). Immobilizationwas carried out at 4˚C with HBS-EP+ as buffer using 200 resonance unitsas a target level for the immobilized ligand. OSCAR was diluted to 25 nMin HBS-EP+ buffer containing 0.1% serum albumin and increasing con-centrations of the anti-OSCAR Abs from 0 to 20 mg/ml. The OSCAR/mAbsolutions were injected for 240 s followed by 200 s dissociation at a flowrate of 50 ml/min at 4˚C followed by regeneration with 10 mM glycine (pH2.0). Binding curves were double referenced by subtraction of the signalsfrom reference cells and a buffer injection and further analyzed using theBiacore T100 evaluation software.

Cells

Buffy coats from healthy individuals were obtained anonymously from theBlood Bank, Rigshospitalet, Copenhagen, Denmark. All donors gave in-formed consent according to the protocol approved by the Ethics Committeefor Copenhagen, Denmark, for research use (approval no. H-D-2008-113).PBMCs were freshly purified by Ficoll-Plaque density centrifugation.Monocytes were purified from PBMCs using the CD14MicroBeads positiveselection kit (Miltenyi Biotec). To induce DC differentiation, monocyteswere cultured in RPMI 1640/10% FCS in the presence of 50 ng/mlrecombinant human GM-CSF and 50 ng/ml recombinant human IL-4(Novo Nordisk) for 6 d. These cells are referred to as immature DCs(iDCs). Naive CD4+RA+ cells were purified from PBMCs using a humannaive CD4+ T cell enrichment kit (StemCell Technologies). For freezing,T cells were resuspended at a density of 7 3 106/ml in RPMI 1640/50%FBS/10% DMSO, and cryovials were placed in a Mr. Frosty container(Thermo Fisher Scientific) at 280˚C overnight and then transferred intoa 2140˚C freezer. Viability of the thawed cells was ∼95% as measuredusing a NucleoCounter (ChemoMetec).

DC culture on collagen

Untreated virgin polystyrene 24-well plates were coated with ColI or ColIIin PBS at ∼3 mg/cm2 overnight at 4˚C. iDCs (13 106) were added per wellafter being preincubated for 20 min with indicated mAb. The cells werecultured for 20 h unless otherwise specified. ColI and ColII had similareffects in all experiments. ColII was always slightly more potent than ColI.Functional data in the manuscript are shown for ColII only unless other-wise specified.

Allogeneic and autologous MLR

For allogeneic MLRs freshly purified cells were used. For autologousreactions, T cells were kept frozen in RPMI 1640/50% FBS/10% DMSOuntil the autologous DCs became matured. T cells (1 3 105) werecocultured in triplicate wells with 1 3 103 DCs in RPMI 1640/10% FCSgrowth medium. No live DCs were detected in the coculture at the time oftermination of the experiment. To assess T cell proliferation pulse labelingof cells with [3H]thymidine was used to monitor DNA synthesis. Briefly, atday 4 of coculture the cells were pulsed with 0.5 mCi [3H]thymidine(PerkinElmer) for 18 h. [3H]thymidine incorporation was measured byliquid scintillation counting using a TopCount NXT (PerkinElmer).

Cytokine secretion

Multiplex analysis of cytokines was performed using the BioPlex 200cytokine assay platform (Bio-Rad). IL-23 (p19/p40) was detected using thePlatinum ELISA kit (eBioscience).

Expression of cell surface markers

For analysis of cell surface marker expression, the following Abs were used:BD multicolor PE-conjugated mouse monoclonal anti-human CD86,PerCP-Cy5.5–conjugated mouse anti-human CD209, allophycocyanin-conjugated mouse anti-human CD83 (BD Biosciences), FITC-conju-gated mouse anti-human HLA-DR (BD Pharmingen), allophycocyanin-conjugated mouse anti-human CD14 (BD Pharmingen), PE-conjugatedmouse anti-human OSCAR (clone 11.1CN5; Beckman Coulter), and Pa-

cific Blue–conjugated anti-human CD3 (BD Pharmingen). Cells werewashed twice in PBS with 2% FCS and then stained for 30 min at 37˚C inPBS with 2% FCS in the presence of 1% human serum albumin to blockunspecific binding. Cells were rinsed three times in PBS with 2% FCSprior to flow cytometry analysis. Data acquisition was done usinga FACSFortessa (BD Biosciences). Only live cells, that is, negative forallophycocyanin-H7 dead cell stain (Invitrogen/Life Technologies) wereanalyzed. For compensation single stain was used with one drop of neg-ative control beads and anti-mouse IgG beads (BD Biosciences). For thedead cell marker compensation, amine reactive compensation beads wereused (Invitrogen/Life Technologies). Data analysis was performed usingKaluza software version 1.2 (Beckman Coulter).

Microarray analysis

Total RNAwas obtained 4 and 20 h after stimulating DCs with ColI/II in theabsence or presence of anti-OSCARmAb or isotype control mAb. RNAwasextracted using TRIzol (Invitrogen/Life Technologies) followed by puri-fication with an RNeasy MinElute cleanup kit (Qiagen). The RNA integritywas evaluated on an Agilent 2100 bioanalyzer using Agilent RNA 6000Nano Kit chips (Agilent Technologies), with RNA integrity number scoresof $9.9. Microarray experiments were performed according to the pro-tocols supplied by Affymetrix and run on the HT HG-U133+ PM chipfrom Affymetrix using an Affymetrix GeneTitan machine in a 96-welllayout. Array data were normalized using the robust multiarray averagealgorithm with a custom chip definition file obtained from BrainArray(University of Michigan) based on Ensembl gene annotations (ENSG).Microarray data were analyzed for significant differences using QlucoreOmics Explorer (v2.2) software. A paired two-group comparison wasperformed with Benjamini–Hochberg correction for multiple testing ap-plying a FDR of 5% and a biological threshold of .2-fold regulation.Differentially expressed genes were subjected to gene ontology analysisfor biological processes using Ingenuity Pathway Analysis. The datasetand technical information compliant with minimum information abouta microarray experiment can be found at the EMBL-EBI ArrayExpressdepository, accession no. E-MTAB-2904 (http://www.ebi.ac.uk/arrayexpress/experiments/E-MTAB-2904/).

Statistical analysis

Statistical analyses were performed in GraphPad Prism 5 using Student two-tailed t tests. A p value ,0.05 was considered statistically significant.Unless otherwise indicated, means and SD are shown.

ResultsIdentification of OSCAR ligand

It was previously shown that a putative ligand for murine OSCAR isexpressed on the cell surface of osteoblasts (18). In the presentstudy, we screened a panel of cell lines for expression of a putativeligand for human OSCAR by monitoring binding of fluorescentlylabeled OSCAR-ECD to the surface of live cells. A murineMC3T3-E1 preosteoblast cell line was selected for purification ofthe OSCAR ligand. Using affinity chromatography followed byliquid chromatography–tandem MS analysis of eluted proteins, wehave identified a1 ColI as OSCAR-interacting protein. Directinteraction of OSCAR with native human ColI an ColII wasconfirmed by SPR, whereas no binding was detected to ColIII orColIV (Supplemental Fig. 1). Thus, ColI and ColII serve as nat-urally occurring OSCAR ligands.Anti-OSCAR mAb capable of blocking OSCAR–collagen in-

teraction as well as nonblocking mAb were selected for furtherfunctional assays using SPR (Supplemental Fig. 2).

OSCAR–collagen signaling promotes DC survival

It has previously been found that cross-linking of OSCAR with anmAb promoted survival of monocyte-derived DCs (11). In thepresent study, we tested whether the naturally occurring OSCARligands, namely ColI/II, could promote DC survival. We differ-entiated iDCs from peripheral blood monocytes in the presence ofGM-CSF and IL-4. This method allows generation of cells phe-notypically resembling “inflammatory” DCs, which differentiatein vivo from blood-derived precursors during inflammation (19).

The Journal of Immunology 3

The iDCs were plated onto either uncoated plates with or withoutGM-CSF, or onto collagen-coated plates without GM-CSF in theabsence or presence of anti-OSCAR–blocking mAb or an isotypecontrol mAb. We assessed cell viability by monitoring the intra-cellular reducing environment (an indicator of cellular well-being)with alamarBlue and by detecting cell death using annexinV/propidium iodine staining. After 7 d of incubation, the reducingpotential was significantly decreased in DCs cultured on plasticwithout GM-CSF and in collagen-coated plates in the presence ofanti-OSCAR mAb. In contrast, the cells maintained the reducingenvironment upon culturing with GM-CSF or on collagen (Fig. 1).Already after 3 d, a large percentage of annexin V+ apoptotic cellswas detected in cultures lacking GM-CSF, whereas most cellscultured on collagen were alive (Table I). The latter effect wascompletely abolished by treatment of cells with anti-OSCARmAb. As control, treatment with Enbrel (etanercept, the p75TNFR-Fc fusion protein), a drug currently used for treatment of RA, wasused, as it has been shown that TNF-a signaling supports DC sur-vival (20). We show that Enbrel treatment partially inhibits collagen-induced survival of DCs. Our data indicate that OSCAR–collagensignaling can promote DC survival under the conditions of growthfactor withdrawal.

Collagen induces secretion of cytokines and chemokines bymonocyte-derived DCs

Cross-linking of OSCAR in iDCs has been shown to induce se-cretion of IL-8, IL-12p40, M-CSF, MCP-1/CCL2, and MDC/CCL22, whereas no secretion of IL-1b, IL-6, IL-10, GM-CSF,IL-12p70, IP-10/CXCL10, and TNF-a was detected (9, 11). Wetested whether collagen can induce secretion of cytokines viaOSCAR signaling. Culturing of iDCs on ColII-coated platescaused a profound induction of multiple cytokines secretion, in-cluding TNF-a, IL-6, IL-8, IL-10, IL-13, IL-23, and RANTES(CCL5). This activation of DCs was completely abolished by anti-OSCAR–blocking mAb (Table II). Meanwhile, anti-OSCAR mAbdid not show any effect on the LPS-stimulated cytokine release(data not shown). In contrast to the positive control, namely LPS-matured DCs, IL-12p70 was not secreted by collagen-matured

DCs (Col-DCs). Thus, OSCAR signaling leads to activation ofDCs characterized by production of a broad range of cytokinesand chemokines, mainly of a proinflammatory type.

Collagen promotes maturation of monocyte-derived DCs

ColI has been shown to induce maturation of monocyte-derivedDCs, but the receptors mediating this effect were not identified(21). Cross-linking of OSCAR with specific Abs induced semi-maturation of DCs (11). We therefore tested whether collagen-induced DC maturation is mediated by OSCAR. iDCs were gen-erated as described above. The phenotype of these cells (CD142

CD209+) was confirmed by flow cytometry. Treatment with anti-OSCAR mAb did not have an effect on differentiation of iDCsunder the conditions used (data not shown). The iDCs were platedonto ColI- (data not shown) or ColII-coated plates and cultured for1–2 d. The cell surface markers were analyzed by flow cytometry(Fig. 2A). We found that collagen drastically upregulated ex-pression of the maturation markers CD83 and CD86. HLA-DRwas upregulated only in donors that had a relatively low basalexpression of this marker (Fig. 2B). This effect was completelyabolished by anti-OSCAR–blocking mAb in a dose-dependentmanner (Fig. 2C), indicating that collagen-induced DC matura-tion is mediated by OSCAR signaling. A mAb against integrina2b1, another known collagen receptor, had no inhibitory effect(data not shown), which is in agreement with the data publishedby Brand et al. (21). LPS, which was used in some experiments aspositive control, showed a level of upregulation of the cell surfacemarkers that was comparable to collagen (data not shown). Mat-uration of DCs can be induced by TNF-a (19). Because we haveshown that collagen induces TNF-a release from DCs, we testedwhether blocking TNF-a would inhibit ColII-induced DC matu-ration. Treatment of the cells with Enbrel only partially inhibitedDC maturation in two of four donors tested (Fig. 2D). Thus,autocrine TNF-a does not seem to be the major contributor to thecollagen-induced DC maturation.

Col-DCs are efficient inducers of T cell proliferation

Triggering of OSCAR with cross-linking mAb was found topotentiate effects of TLR ligands on DC maturation and, asa consequence, to enhance the ability of TLR-matured DCs toinduce proliferation of allogeneic naive T cells. In the absence ofTLR ligands, the DCs treated with anti-OSCAR cross-linkingmAb completely failed to induce T cell proliferation (11). Inthe present study, we tested whether the Col-DCs can promoteproliferation of allogeneic naive CD4+CD45RA+ T cells. TheCol-DCs were very efficient in promoting T cell proliferation asmonitored either by [3H]thymidine incorporation (Fig. 3A) or bythe expression of a proliferation marker, Ki67 (not shown).When the Col-DCs were matured in the presence of anti-OSCARmAb, this proliferative effect was abolished in a dose-dependentmanner.Cytokines secreted by T cells, which had been expanded by Col-

DCs, were analyzed (Table III). Noticeably, a strong induction ofa number of cytokines, including Th2-specific IL-5 and IL-13,as well as proinflammatory IL-6, IL-8, TNF-a, and RANTES(CCL5), was evident in the Col-DC/T cell cocultures, which wasnot the case when DCs were treated with anti-OSCAR mAbduring the maturation step.We further tested the ability of Col-DCs to induce proliferation

of autologous naive T cells. In this setting also, Col-DCs wereefficient inducers of T cell proliferation (Fig. 3B). The secretion ofcytokines was analyzed as above (Table IV). We detected highlevels of IL-13 and IL-5, the cytokines that polarize T cells towardthe Th2 lineage. However, the proinflammatory cytokines IL-6,

FIGURE 1. OSCAR–collagen signaling promotes DC survival. Human

monocytes were purified and differentiated into iDCs by GM-CSF and IL-4

(50 ng/ml each) for 6 d. Then, 1 3 106 iDCs/ml were cultured on either

uncoated or ColII-coated plates. To the ColII-coated wells, OSCAR-

blocking Ab (1 mg/ml) or control Ab (1 mg/ml) was added. GM-CSF (50

ng/ml) was added as a positive control of survival. As a negative control,

iDCs were left in growth factor–free medium. After 7 d, cell viability was

assessed using alamarBlue. Data are expressed as mean of biological trip-

licates 6 SD. Data shown are from one representative donor out of five

analyzed.

4 OSCAR-DEPENDENT MATURATION OF DCs

IL-8, and TNF-a as well as Th1-polarizing INF-g were also in-duced in an OSCAR-dependent manner. Thus, no clear pattern ofT cell polarization was evident.

Signaling by OSCAR–collagen regulates gene expression inDCs

The overall gene expression profile resulting from the OSCAR–collagen interaction in DCs was characterized by Affymetrix genechip analysis. iDCs were exposed to ColI or ColII for 4 or 20 h inthe presence or absence of antagonistic anti-OSCAR mAb orisotype control mAb. The effects of ColI and ColII were verysimilar (.90% of genes regulated by ColI were also regulated byColII). Only data for the 4-h ColII treatment are discussed in thissection, as it more closely represents the primary response toOSCAR stimulation without following amplification (secondaryresponse). LPS was included as a positive control for DC matu-ration, and there was a good overlap (∼70%) in gene expressionchanges in cells treated with LPS and ColI/II. Raw data files for alltreatments can be downloaded from the ArrayExpress depository(accession no. E-MTAB-2904).A marked change in the gene expression level was observed

for 709 genes in cells treated with ColII for 4 h (SupplementalTable IA). When comparing the effect of preincubating the iDCswith anti-OSCAR mAb or isotype control mAb prior to plating onColII, the expression level of 546 genes was significantly differentbetween the two treatments (Supplemental Table IB). No signif-icant differences (FDR of 10%) were found for treatments withColII in the presence or absence of isotype mAb, negating an effectof the isotype control mAb on OSCAR biology.When comparing the ColII-regulated genes and genes regulated

in the presence of anti-OSCAR mAb, 472 genes were overlapping.Thus, these genes are thought to be the true transcriptional tar-

gets of the OSCAR–collagen signaling pathway (SupplementalTable IC). Using Ingenuity Pathway Analysis, the OSCAR-regulated genes were clustered into several different functionalpathways. The top 20 canonical pathways are listed in Table V.The OSCAR-dependent genes were mainly associated with im-mune regulation, maturation, adhesion, apoptosis, and varioussignal transduction pathways. The data show that OSCAR reg-ulates expression of the major components of the stable immunesynapse such as CD40, CD80, CD86, CD83, and ICAM-1. Fur-thermore, OSCAR signaling induced proinflammatory cytokineand chemokine gene expression as well as the genes encodingcytokine and chemokine receptors. The gene encoding CCL20,a chemokine attracting iDCs, was noticeably the most upregulatedgene with a fold change of .150. Of note, chemoattractants forprecursors of endothelial cell CXCL1, -2, and -3 mRNA levelswere drastically induced (.25-fold) at a transcriptional level(Supplemental Table IC), indicating possible involvement of theOSCAR pathway in the regulation of neoangiogenesis. OSCAR–collagen signaling upregulated genes encoding components ofTLR, TREM-1, and uPA/MMP-9 pathways highly relevant forpathogenesis of RA and other autoimmune diseases (22). Of in-terest is also the downregulation of the Wnt/fzd pathway thatprograms DCs to induce tolerogenic responses (23). We observedthat ColII decreased FZD2 gene expression by 18-fold, which wasabrogated by treatment with anti-OSCAR mAb. Taken together,our data indicate that OSCAR signaling in DCs has a profoundeffect on the induction/maintenance of proinflammatory responsesat multiple levels.

DiscussionUntil recently, the naturally occurring ligand for OSCAR remainedunknown. In the present study we have identified OSCAR ligand

Table I. DC viability analyzed by flow cytometry

Alive (PI2/Annexin V2) Early Apoptotic (PI2/Annexin V+) Necrotic (PI+/Annexin V2) Late Apoptotic (PI+/Annexin V+)

Untreated 18.1 6 1.3 30.6 6 0.5 11.8 6 1.2 39.7 6 1.9GM-CSF 65.0 6 5.6** 17.4 6 1.7** 2.7 6 1.3* 14.9 6 3.0 ***ColII 83.9 6 1.4*** 9.1 6 0.3*** 0.4 6 0.1** 6.6 6 1.1**ColII + anti-OSCAR 13.4 6 1.1* 31.0 6 0.3 14.2 6 3.5 41.5 6 4.1ColII + isotype 80.6 6 3.8** 10.4 6 1.6** 0.3 6 0.1** 8.7 6 2.1**ColII + Enbrel 59.3 6 4.4** 20.7 6 2.0* 0.8 6 0.1** 19.1 6 2.6*

iDCs were cultured on ColII-coated plates. Cells were untreated or treated with either 50 ng/ml GM-CSF, 1 mg/ml anti-OSCAR, 1 mg/ml isotype control, or 10 mg/ml Enbrel.After 3 d of culture cells were stained for annexin Vand propidium iodide (PI) and analyzed via flow cytometry. For flow cytometry analysis, single cells were gated on with sidescatter (SSC) height versus SSC width, followed by SSC versus forward scatter where debris was gated out. The percentages of cell populations were analyzed based on a dotplot (annexin V versus PI). Data are expressed as means 6 SD of triplicate wells treated independently. Data are shown for one representative donor out of four analyzed.

*p , 0.05, **p , 0.01, ***p , 0.001 versus untreated.

Table II. Cytokines and chemokines secreted by DCs

iDCs Col-DCs Col-DCs + Anti-OSCAR Col-DCs + Isotype mAb LPS-DCs

IL-6 2,428.0 6 1.5 8,565.0 6 175.8* 1,729.0 6 97.7** 6,750.0 6 501.1ns 22,547.0 6 364.4**IL-8 11,406.0 6 485.5 20,167.0 6 523.4** 11,809.0 6 457.6** 19,911.0 6 699 ns 20,286.0 6 508.0**IL-10 239.5 6 48.4 2,801.0 6 188.0* 85.5 6 11.0* 2,330.0 6 48.4ns 12,049.0 6 126.4**IL-12p40 536.2 6 35.0 46,361.0 6 406.0** 595.6 6 49.0** 46,937.0 6 2,298.0ns 39,407.0 6 1771.0*IL-12p70 71.5 6 33.7 99.7 6 6.1ns 40.1 6 1.5* 86.7 6 0.0ns 14,287.0 6 315.8**IL-13 21.2 6 11.2 1,034.0 6 34.3* 11.5 6 0.0* 1,039.0 6 34.7ns 1,658.0 6 6.9**RANTES 64.0 6 15.4 2,201.0 6 132.1* 54.8 6 0.5* 2,131.0 6 52.9ns 44,79.0 6 193.4*TNF-a 366.6 6 101.2 44,763.0 6 1881.0* 78.1 6 0.0* 28,719.0 6 2669.0ns 31,108.0 6 1961.0*IL-23 0.4 6 0.4 272.2 6 6.5*** 0.8 6 3.4*** 289.3 6 40.1ns 696.3 6 8.6***

Multiple cytokines were measured in the supernatants from DCs cultured at the indicated conditions using Luminex technology. Values are shown in pg/ml. Data representmeans 6 SD of triplicate wells treated independently. Data are shown for 1 representative donor out of .20 donors analyzed with similar results. IL-23 was measured usinga sandwich ELISA.

nsnonsignificant.LPS-DCs, LPS-matured DCs.*p , 0.05, **p , 0.01, ***p , 0.001 by t test for the following conditions: iDCs versus Col-DCs and for Col-DCs versus Col-DCs plus anti-OSCAR.

The Journal of Immunology 5

FIGURE 2. Expression of cell surface markers on DCs analyzed by flow cytometry. Initially, 1 3 106 iDCs/ml were cultured on either uncoated or ColII-

coated plates for 18 h. At these conditions, DCs upregulated expression of maturation markers when cultured on collagen. (A) Top panel, Gating strategy for

flow cytometry analysis. Only cells negative for the dead cell marker Near-IR were used in the analysis (gate a). Single cells from gate a were further gated on

with side scatter height (SSC-H) versus SSC width (SSC-W) (gate b), followed by gating out cell debris with SSC versus forward scatter (FSC) (gate c).

Hierarchy of gating is indicated with arrows. Only cells from gate c were further analyzed. Lower panel, Representative histograms are shown for the in-

dicated surface markers. Open histograms represent unstained cells. Gray and black histograms represent iDCs and Col-DCs, respectively. (B) Combined data

are shown for 16 donors. (C) ColII-induced expression of maturation markers is downregulated by anti-OSCAR mAb in a dose-dependent manner; 0.01, 0.1,

and 1 mg/ml anti-OSCAR mAb or 1 mg/ml isotype control mAb was used. Data are expressed as mean of independently treated triplicate wells 6 SD. Data

are shown for 1 representative donor out of 10. (D) DCs treated with 10 mg/ml Enbrel or 1 mg/ml isotype control mAb. Data for four donors are shown.

6 OSCAR-DEPENDENT MATURATION OF DCs

expressed on murine preosteoblasts as being a1 ColI. OSCARinteraction with native human ColI and ColII was confirmed bythe SPR technique. While we were conducting our experiments,the Trowsdale et al. (14) published their discovery of collagens asligands of OSCAR. In this study, the authors used a syntheticColII peptide library and identified a number of peptides showingdirect binding to human OSCAR. Some of these peptides wereefficient stimulators of receptor activator for NF-kB ligand–de-pendent osteoclastogenesis. The authors confirmed binding ofOSCAR to native human ColI, ColII, and ColIII. We neither sawa binding to native ColIII using SPR analysis nor a biologicaleffect on DCs, which might be attributed to differences in qualitiesof the native protein preparations used in the two studies.Collagens are broadly distributed in the human and, in addition

to their structural function, they play an important role in celladhesion, migration, proliferation, differentiation, and survival(24). These multiple and diverse effects are highly dependent onthe ECM structure and remodeling (25). ColI is the main proteinconstituent of bone, whereas ColII is the major protein componentof cartilage. Under pathological conditions, such as, for example,RA, the ECM undergoes excessive remodeling with drastic up-regulation of the ColI–III turnover within the joints (26). Wesuggest that under such conditions, collagens of bone and cartilagemay become available for interaction with infiltrating immunecells and may considerably influence their activation status andviability within RA joints. Indeed, it has been shown that OSCAR-expressing mononuclear cells are in contact with collagens in thejoints and the OSCAR-binding sites are expected to be on thesurface of collagens exposed to the infiltrating cells (14).Among the cells expressing the OSCAR protein and potentially

capable of entering synovial tissue upon inflammation aremonocytes, monocyte-derived DCs, and lineage-negative CD11c+

DC (9). Whereas the latter are thought to represent a rare het-erogeneous population of cells found in health and disease,monocytes may serve as precursors of inflammatory DCs, whichare only generated in pathological conditions, such as, for ex-ample, RA (27). Development of such inflammatory, fully func-tional DCs is commonly modeled in vitro by treatment ofperipheral blood monocytes with GM-CSF and IL-4 (19). Themonocyte-derived DCs are potent APCs, which secrete multipleproinflammatory cytokines and express CD1a, a surface marker

that was shown to be specific for this cell type (28). CD1a+ cellswere found in RA synovial tissue, where they are thought to befunctional (29). In the present study, we chose to focus on thispopulation of inflammatory DCs to mechanistically address theoutcome of OSCAR–collagen signaling with relevance to the RApathogenesis, although we cannot exclude an impact of theOSCAR–collagen interaction on the Ag-presenting function insubpopulations of the CD11c+ cells, which should further beaddressed in clinical studies.Merck et al. (9, 11) have shown that cross-linking of OSCAR

with the receptor-specific Abs can provide survival, activation, andmaturation signals to human monocyte-derived DCs. The use ofcross-linking mAbs to stimulate the ITAM-bearing receptors viaaggregation is an efficient way to address the signaling potential ofa receptor when its natural ligand is unknown. However, an en-dogenous ligand may induce a signal of different magnitude andquality, depending on the affinity of interaction and cross-talkingsignaling pathways modulating the ITAM signal. Therefore, it isessential to investigate the outcome of the OSCAR signaling in thepresence of the native ligand in the cells where other collagenreceptors are coexpressed.OSCAR expression in human monocyte-derived DCs (11) and

monocytes (30) was suggested to support cell survival in the ab-sence of growth factors via induction of antiapoptotic moleculesBcl-2 and Bcl-xL through PI3K and ERK signaling. In our study,the DCs plated on ColI/II showed prolonged viability in the ab-sence of the essential growth factor, GM-CSF, indicating thatOSCAR could play an important role in cell survival. Expressionof genes encoding multiple components of the ERK pathway(including MAP3K1 and MAP3K8), as well as BCL2A1 encodinga Bcl-2 related protein, was strongly upregulated by the OSCAR–Col signaling. Autocrine TNF-a was previously shown to providesurvival signals to DCs (20). Blocking of TNF-a with eithersoluble p55TNFR-Fc or anti–TNF-a mAb (infliximab) signifi-cantly reduced survival of LPS-matured DCs (31). We show in thepresent study that survival of Col-DCs was impaired by treatmentwith Enbrel, although to a much lower degree than by treatmentwith anti-OSCAR mAb. These data indicate that collagen supportssurvival of DCs at least in part via an autocrine TNF-a signaling.DCs plated on ColI/II secreted a number of cytokines with

a proinflammatory function. In contrast to the OSCAR cross-

FIGURE 3. ColII-matured DCs induce proliferation of T cells. (A) Allogeneic MLR. Proliferation of CD3+CD4+CD45+RA+ T cells was estimated by

[3H]thymidine incorporation. T cells were incubated for 4 d with the indicated allogeneic DCs: iDCs, ColII-matured DCs, or DCs that were maintained on

collagen in the presence of anti-OSCAR mAb (data shown for 0.01, 0.1, and 1 mg/ml mAb) or isotype control mAb (1 mg/ml), as well as T cells alone (no

DCs) as negative control and T cells stimulated with CD3/CD28 beads as positive control. (B) Proliferation of autologous CD3+CD4+CD45+RA+ T cells in

the presence of the indicated DCs; costimulation of TCR was achieved by treatment with an agonistic OKT-3 mAb; all treatments and controls are as in (A).

Data are mean 6 SD of triplicate wells from one representative experiment out of seven (allogeneic) or three (autologous) performed with similar results.

The Journal of Immunology 7

linking data (9, 11), we observed a profound induction of IL-6, IL-1b, and TNF-a by ColI/II. The discrepancy might be due toa higher sensitivity of the detection method used in the presentstudy (Luminex versus ELISA) or due to the differences in bio-logical set-up, where exposure to ColI/II may provide additionalsignals necessary for more efficient performance of the OSCARpathway. Most of the cytokines and their receptors were found tobe transcriptional targets of OSCAR based on the microarrayanalysis. A number of chemokines were also induced by OSCAR–collagen signaling, including CCL20 (recruiting iDCs), IL-8(chemoattractant for granulocytes), and RANTES (CCL5, che-moattractant for T cells) as well as attractants for endothelial cellprecursors CXCL1, -2, and -3. The components of the pathwaysregulating expression of these proinflammatory mediators (e.g.,

NF-kB pathway members, STAT pathway members, MAPKs)were also upregulated by collagen in an OSCAR-dependentmanner at the transcriptional level. Merck et al. (11) haveshown potentiation of TLR signaling by OSCAR cross-linkingcharacterized by secretion of proinflammatory cytokines andmaturation of DCs. In the present study, we provide evidence thatthe OSCAR signaling pathway upregulates transcription of thegenes encoding receptors and multiple components of the TLRsignaling machinery.Earlier studies indicated that ColI can stimulate maturation of

human monocyte-derived DCs in an integrin-independent manner(21), the mechanism for which so far remained undiscovered. Ourdata show that the OSCAR signaling pathway controls Col-induced DC maturation. The iDCs plated on ColI/II showed fast

Table III. Cytokines and chemokines secreted by T cells cocultured with allogeneic DCs

iDCs Col-DCsCol-DCs +Anti-OSCAR

Col-DCs +Isotype mAb

IL-2 38.4 6 13.2 3,135.0 6 463.0** 35.8 6 18.0** 2,666.0 6 543.0ns

IL-3 811.8 6 16.4 771.7 6 6.1* 782.3 6 62.0ns 751.7 6 65.5ns

IL-5 1.4 6 0.2 2,087.0 6 234.0** 1.3 6 0.0** 1,919.0 6 598.9ns

IL-6 12.0 6 3.7 1,414.0 6 329.5* 20.0 6 10.8* 1,569.0 6 163.7ns

IL-8 1,035.0 6 761.8 10,852.0 6 859.8** 1,352.0 6 458.4** 10,936.0 6 274.1ns

IL-10 13.9 6 1.4 99.0 6 65.2ns 14.2 6 2.5ns 147.2 6 54.8ns

IL-12p40 1,339.0 6 65.4 10,083.0 6 292.1** 1,349.0 6 56.5*** 13,601.0 6 403.8ns

IL-12p70 10.4 6 0.0 13.0 6 0.0ns 11.3 6 1.5ns 14.3 6 1.3ns

IL-13 2.5 6 0.8 3,547.0 6 142.3*** 2.5 6 0.2*** 3,609.0 6 321.5ns

IL-17 46.1 6 3.2 48.9 6 3.2ns 48.9 6 3.2ns 49.4 6 5.6ns

IL-18 25.8 6 2.5 25.2 6 1.4ns 24.2 6 0.7ns 26.4 6 1.6ns

GM-CSF 26.1 6 3.9 1,402.0 6 217.4** 23.9 6 2.5** 1,581.0 6 135.2ns

IFN-g 24.5 6 2.0 254.9 6 69.4* 23.1 6 2.4* 514.0 6 195.1ns

MCP-1 992.4 6 549.3 85.6 6 2.3ns 1,505.0 6 654.3ns 86.0 6 8.8ns

M-CSF 268.5 6 12.0 347.6 6 20.0* 248.4 6 6.4** 357.1 6 30.7ns

RANTES 14.6 6 6.3 2,180.0 6 230.4** 18.4 6 9.6** 2,287.0 6 152.2ns

TNF-a 104.3 6 56.4 3,830.0 6 168.4*** 101.9 6 18.1*** 3,575.0 6 567.1ns

Multiple cytokines were measured in the supernatants of T cells cultured with the indicated allogeneic DCs. Supernatantswere collected at day 4 of coculture. Anti-OSCAR mAbs or isotype control mAbs were used at 1 mg/ml. Values are shown inpg/ml. Data represent means 6 SD of triplicate wells treated independently. Data are shown for one representative experimentout of four independent experiments with similar results.

nsnonsignificant.p , 0.05, **p , 0.01, ***p , 0.001 by t test for the following conditions: iDCs versus Col-DCs, Col-DCs versus Col-

DCs + anti-OSCAR, and Col-DCs versus Col-DCs + isotype.

Table IV. Cytokines and chemokines secreted by T cells cocultured with autologous DCs

iDCs Col-DCs Col-DCs + Anti-OSCAR Col-DCs + Isotype mAb

IL-2 16.2 6 0.9 18.0 6 0.6ns 14.0 6 1.7ns 15.6 6 0.6ns

IL-3 2397.0 6 140.9 2,399.0 6 85.2ns 2,225.0 6 186.1ns 2,456.0 6 47.0ns

IL-5 52.8 6 23.9 377.9 6 67.0* 45.3 6 22.1* 302.7 6 19.4ns

IL-6 894.0 6 97.4 2,595.0 6 202.3** 395.1 6 88.3** 3,052.0 6 1,125.0ns

IL-8 2,1579.0 6 17.9 22,683.0 6 270.7* 18,465.0 6 1,058* 21,496.0 6 328.1ns

IL-10 213.8 6 9.0 430.7 6 10.6** 188.2 6 23.1** 419.7 6 16.1ns

IL-12p40 20,956.0 6 3,700.0 38,572.0 6 732.8* 10,840.0 6 1,641** 43,733.0 6 3518ns

IL-12p70 28.9 6 2.7 19.3 6 0.0ns 16.4 6 1.4ns 20.2 6 2.7ns

IL-13 207.4 6 0.2 966.1 6 28.3*** 158.5 6 18.4*** 884.3 6 114.4ns

IL-17 52.3 6 5.5 67.8 6 11.0ns 42.1 6 4.8ns 55.2 6 1.4ns

IL-18 20.5 6 2.5 21.0 6 0.8ns 20.5 6 0.8ns 21.6 6 2.5ns

GM-CSF 227.5 6 4.2 495.0 6 58.6* 230.3 6 11.0* 472.3 6 83.2ns

IFN-g 1,429.0 6 352.4 3,236.0 6 954.6ns 213.3 6 166.2* 3,704.0 6 1,330.0ns

MCP-1 234.7 6 52.8 8.0 6 1.1* 372.9 6 35.4** 8.1 6 1.8ns

M-CSF 893.7 6 6.0 1,873.0 6 82.6** 908.0 6 38.3** 2,163.0 6 3.0*RANTES 1,416.0 6 0.9 1,262.0 6 85.3ns 2,408.0 6 63.6** 15,12.0 6 182.8ns

TNF-a 1,706.0 6 305.0 8,297.0 6 507.3** 2,016.0 6 377.2** 9251.0 6 855.6ns

Multiple cytokines were measured in the supernatants of T cells cultured with the indicated autologous DCs. Supernatants were collected at day 4 of coculture. Anti-OSCARmAbs or isotype control mAbs were used at 1 mg/ml. Values are shown in pg/ml. Data represent means 6 SD of triplicate wells treated independently. Data are shown for onerepresentative experiment out of four independent experiments with similar results.

nsnonsignificant.*p , 0.05, **p , 0.01, ***p , 0.001 by t test for the following conditions: iDCs versus Col-DCs, Col-DCs versus Col-DCs + anti-OSCAR, and Col-DCs versus Col-DCs +

isotype.

8 OSCAR-DEPENDENT MATURATION OF DCs

and drastic upregulation of maturation markers. Microarrayanalysis has shown that most of the major components of thestable immune synapse, including CD40, CD80, CD86, andICAM-1, are transcriptional targets of the OSCAR–collagen sig-naling. Treatment of cells with anti-OSCAR mAb completelyblocked this collagen-induced maturation process. The cellsurface expression of HLA-DR was upregulated in an OSCAR-dependent manner. However, expression of the MHC class IIgenes was not influenced by collagen exposure, indicating reg-ulation at a posttranscriptional level. These results are in linewith the previously published studies showing posttranscrip-tional regulation of the surface expression of MHC class II (32),particularly via protein recycling/trafficking to and from theplasma membrane (33, 34).

In contrast to the OSCAR cross-linking data (11), in our studytriggering of the OSCAR signaling in DCs with a native ligand ledto formation of matured cells, which efficiently stimulated pro-liferation of allogeneic and autologous naive T cells. Thus, even inthe absence of microbial or inflammatory stimulation, OSCARengagement by collagen was sufficient to induce a full DC mat-uration program. The T cells expanded by Col-DCs producedmultiple cytokines mainly of the proinflammatory type, but withno clear signs of T cell polarization. Additional studies on phe-notyping of these cells (e.g., microarray and transcription factorprofiling) are needed to fully understand the T cell polarizingeffect of the Col-DCs.Upregulation of OSCAR expression on the surface of blood

monocytes was found in RA patients with active disease, and

Table V. Top 20 canonical pathways regulated by ColII and reversed by anti-OSCAR

Ingenuity Canonical Pathways Molecules 2Log (p Value)

TLR signaling MAP3K14, IL1A, MAP3K1, TLR8, TNFAIP3, MAPK13, NFKB1,TLR2, NFKBIA, IL1RN, IL1B, MAP2K3, TNF, TRAF1, IRAK2

1.05E+01

Hepatic fibrosis/hepatic stellate cell activation IL4R, IL1A, ICA, VEGFA, CXCL3, CCL2, CD40, EDN1, CSF1,TNFSF8, IL1B, SERPINE1, IL1RAP, TNF, MMP9, TNFSF14

1.04E+01

Atherosclerosis signaling IL1A, ICAM1, PDGFA, NFKB1, F3, TNFRSF12A, SELPLG,CD40, CCL2, CSF1, IL1RN, LPL, IL1B, SERPINA1, TNF, MMP9,

TNFSF14

9.06E+00

Type I diabetes mellitus signaling SOCS3, MAP3K14, IFNGR1, MAPK13, MAP3K5, JAK2, NFKB1,NFKBIA, CD80, IL1B, CD86, BID, MAP2K3, CASP8, IL1RAP,

TNF

8.89E+00

Role of macrophages, fibroblasts, andendothelial cells in RA

SOCS3, IL1A, ICAM1, PDGFA, TLR8, JAK2, NFKB1, VEGFA,NFAT5, NFKBIA, CCL2, OSM, MAPKAPK2, IL1RAP, FZD2,

TRAF1, MAP3K14, C5AR1, CREB3, TLR2, CSF1, IL1RN, IL1B,MAP2K3, TNF, IRAK2

8.86E+00

Acute phase response signaling SOCS3, MAP3K14, IL1A, C3, MAP3K1, JAK2, MAPK13,MAP3K5, NFKB1, NFKBIA, SOD2, IL1RN, OSM, IL1B,

SERPINA1, MAP2K3, SERPINE1, IL1RAP, TNF

8.48E+00

Granulocyte adhesion and diapedesis IL1A, C5AR1, ICAM1, CCL20, CCL22, SDC4, SELPLG, CXCL3,CLDN12, CCL2, CLDN1, IL1RN, IL1B, CXCL1, CXCL2,

IL1RAP, TNF, MMP9, CCL1

8.14E+00

IL-6 signaling SOCS3, MAP3K14, IL1A, MAPK13, JAK2, NFKB1, VEGFA,NFKBIA, IL1RN, IL1B, MAP2K3, MAPKAPK2, IL1RAP, TNF,

MCL1

7.71E+00

IL-10 signaling SOCS3, MAP3K14, IL1A, IL4R, NFKBIA, IL1RN, IL1B,MAP2K3, MAPK13, NFKB1, TNF, IL1RAP

7.54E+00

Altered T cell and B cell signaling in RA MAP3K14, IL1A, SLAMF1, TLR8, NFKB1, TLR2, CD80, CD40,CSF1, IL1RN, CD86, IL1B, TNF

7.46E+00

TREM1 signaling TLR2, CXCL3, TREM1, ICAM1, CCL2, CD40, TLR8, CD86,IL1B, JAK2, NFKB1, TNF

7.27E+00

CD40 signaling MAP3K14, ICAM1, NFKBIA, CD40, TNFAIP3, MAP2K3,MAPK13, PTGS2, MAPKAPK2, NFKB1, TRAF1

7.03E+00

CD27 signaling in lymphocytes MAP3K14, NFKBIA, MAP3K1, BID, MAP3K8, MAP2K3,MAP3K5, CASP8, NFKB1, MAP3K2

6.99E+00

Agranulocyte adhesion and diapedesis IL1A, C5AR1, ICAM1, CCL20, CCL22, SDC4, SELPLG, CXCL3,CLDN12, CCL2, CLDN1, IL1RN, IL1B, CXCL1, CXCL2, TNF,

MMP9, CCL1

6.98E+00

NF-kB signaling MAP3K14, IL1A, FLT1, MAP3K1, TLR8, TNFAIP3, MALT1,NFKB1, TGFBR2, TLR2, NFKBIA, CD40, IL1RN, IL1B,

MAP3K8, CASP8, TNF

6.86E+00

Role of IL-17A in arthritis CXCL3, NFKBIA, CCL2, CCL20, MAP2K3, CXCL1, MAPK13,PTGS2, MAPKAPK2, NFKB1

6.67E+00

DC maturation MAP3K14, IL1A, ICAM1, CREB3, MAPK13, JAK2, NFKB1,STAT4, TLR2, NFKBIA, CD40, CD80, IL1RN, CD86, IL1B, IRF8,

TNF

6.64E+00

BCR signaling BLNK, MAP3K14, CREB3, MAP3K1, MAPK13, MALT1,MAP3K5, NFKB1, INPP5D, NFKBIA, NFAT5, SYK, MAP2K3,

MAP3K8, RASSF5, BCL2A1, MAP3K2

6.57E+00

TWEAK signaling CASP6, MAP3K14, NFKBIA, BID, CASP8, NFKB1,TNFRSF12A, TRAF1

6.42E+00

TNFR1 signaling CASP6, MAP3K14, NFKBIA, MAP3K1, TNFAIP3, BID, CASP8,NFKB1, TNF

6.17E+00

iDCs were cultured on ColII for 4 h in the presence or absence of GM-CSF (50 ng/ml), anti-OSCAR mAb (1 mg/ml), or isotype control (1 mg/ml). Total cellular RNA wasisolated, and genome-wide mRNA levels were determined using Affymetrix GeneChip microarray analysis. The table shows canonical pathways for OSCAR-ColII–regulatedgenes at an FDR of 5% and a biological criterion of minimum 2-fold change. A complete list of all genes regulated by the OSCAR-ColII signaling at 4 h time point is provided inSupplemental Table IC. Samples from six independent experiments were analyzed.

The Journal of Immunology 9

OSCAR-expressing mononuclear cells were detected in peri-vascular areas of the synovium (13). The RA synovium is con-sidered an ectopic lymphoid organ (35). Both immature andmature DCs are present in the RA synovium in about a 1:1 ratio.The immature cells are primarily located in the sublining or lininglayer of the synovium and in the perivascular infiltrates, whereasmature DCs are associated with perivascular lymphocytic infil-trates (36, 37). These perivascular DCs were suggested to origi-nate from blood-derived precursors that migrated through theactivated endothelium and received differentiation signals, such asGM-CSF and IL-13, within the joint (reviewed in Ref. 38). It isthought that functional DC/T cell interaction can take place lo-cally (39, 40), and the perivascular region may act as a site for cellactivation (41). Intriguingly, surgical removal of cartilage in RApatients led to reduced inflammation and suppression of functionaldifferentiation of DCs in the synovium (42). Because ColII makesup to 60% of dry weight of articular cartilage (43), we speculatethat this observed effect might at least partially be attributed tointerference with the OSCAR signaling in DCs.Taken together, our data indicate that the principal role of

OSCAR in DCs is proinflammatory. Although OSCAR is expressedin resting myeloid cells (9, 13), it may not be functional becausethe ligand accessibility is limited. We hypothesize that theOSCAR–collagen signaling plays a pivotal role in pathology, such asin RA, where ligands become available due to ongoing tissueremodeling. OSCAR may act at multiple levels contributing toperpetuation of local synovial inflammation and may impactsystemic inflammation in RA via expansion of proinflammatoryT cells.

AcknowledgmentsWe thank Lise F.W. Jensen, JetteW. Platou, JingWang, Bing Zhang, Helene

Bregnbak, Jian Zhou, Wenjuan Xia, Xiaoai Wu, Li Yang, and Chunyan Ma

for excellent technical assistance.

DisclosuresThe authors have no financial conflicts of interest.

References1. McInnes, I. B., and G. Schett. 2011. The pathogenesis of rheumatoid arthritis. N.

Engl. J. Med. 365: 2205–2219.2. Holmdahl, R., V. Malmstrom, and H. Burkhardt. 2014. Autoimmune priming,

tissue attack and chronic inflammation—the three stages of rheumatoid arthritis.Eur. J. Immunol. 44: 1593–1599.

3. Thomas, R., L. S. Davis, and P. E. Lipsky. 1994. Rheumatoid synovium is enrichedin mature antigen-presenting dendritic cells. J. Immunol. 152: 2613–2623.

4. Radstake, T. R., A. W. van Lieshout, P. L. van Riel, W. B. van den Berg, andG. J. Adema. 2005. Dendritic cells, Fcg receptors, and Toll-like receptors: po-tential allies in the battle against rheumatoid arthritis. Ann. Rheum. Dis. 64:1532–1538.

5. Lebre, M. C., and P. P. Tak. 2008. Dendritic cell subsets: their roles in rheu-matoid arthritis. Acta Reumatol. Port. 33: 35–45.

6. Radstake, T. R., R. van der Voort, M. ten Brummelhuis, M. de Waal Malefijt,M. Looman, C. G. Figdor, W. B. van den Berg, P. Barrera, and G. J. Adema.2005. Increased expression of CCL18, CCL19, and CCL17 by dendritic cellsfrom patients with rheumatoid arthritis, and regulation by Fcg receptors. Ann.Rheum. Dis. 64: 359–367.

7. Rivollier, A., M. Mazzorana, J. Tebib, M. Piperno, T. Aitsiselmi, C. Rabourdin-Combe, P. Jurdic, and C. Servet-Delprat. 2004. Immature dendritic cell trans-differentiation into osteoclasts: a novel pathway sustained by the rheumatoidarthritis microenvironment. Blood 104: 4029–4037.

8. Alnaeeli, M., J. M. Penninger, and Y. T. Teng. 2006. Immune interactions withCD4+ T cells promote the development of functional osteoclasts from murineCD11c+ dendritic cells. J. Immunol. 177: 3314–3326.

9. Merck, E., C. Gaillard, D. M. Gorman, F. Montero-Julian, I. Durand,S. M. Zurawski, C. Menetrier-Caux, G. Carra, S. Lebecque, G. Trinchieri, andE. E. Bates. 2004. OSCAR is an FcRg-associated receptor that is expressed bymyeloid cells and is involved in antigen presentation and activation of humandendritic cells. Blood 104: 1386–1395.

10. Ishikawa, S., N. Arase, T. Suenaga, Y. Saita, M. Noda, T. Kuriyama, H. Arase,and T. Saito. 2004. Involvement of FcRg in signal transduction of osteoclast-associated receptor (OSCAR). Int. Immunol. 16: 1019–1025.

11. Merck, E., B. de Saint-Vis, M. Scuiller, C. Gaillard, C. Caux, G. Trinchieri, andE. E. Bates. 2005. Fc receptor g-chain activation via hOSCAR induces survivaland maturation of dendritic cells and modulates Toll-like receptor responses.Blood 105: 3623–3632.

12. Tenca, C., A. Merlo, E. Merck, E. E. Bates, D. Saverino, R. Simone, D. Zarcone,G. Trinchieri, C. E. Grossi, and E. Ciccone. 2005. CD85j (leukocyte Ig-likereceptor-1/Ig-like transcript 2) inhibits human osteoclast-associated receptor-mediated activation of human dendritic cells. J. Immunol. 174: 6757–6763.

13. Herman, S., R. B. M€uller, G. Kronke, J. Zwerina, K. Redlich, A. J. Hueber,H. Gelse, E. Neumann, U. M€uller-Ladner, and G. Schett. 2008. Induction ofosteoclast-associated receptor, a key osteoclast costimulation molecule, inrheumatoid arthritis. Arthritis Rheum. 58: 3041–3050.

14. Barrow, A. D., N. Raynal, T. L. Andersen, D. A. Slatter, D. Bihan, N. Pugh,M. Cella, T. Kim, J. Rho, T. Negishi-Koga, et al. 2011. OSCAR is a collagenreceptor that costimulates osteoclastogenesis in DAP12-deficient humans andmice. J. Clin. Invest. 121: 3505–3516.

15. Canfield, S. M., and S. L. Morrison. 1991. The binding affinity of human IgG forits high affinity Fc receptor is determined by multiple amino acids in the CH2domain and is modulated by the hinge region. J. Exp. Med. 173: 1483–1491.

16. Schneider, C., R. A. Newman, D. R. Sutherland, U. Asser, and M. F. Greaves.1982. A one-step purification of membrane proteins using a high efficiencyimmunomatrix. J. Biol. Chem. 257: 10766–10769.

17. Shevchenko, A., M. Wilm, O. Vorm, and M. Mann. 1996. Mass spectrometric se-quencing of proteins silver-stained polyacrylamide gels. Anal. Chem. 68: 850–858.

18. Kim, N., M. Takami, J. Rho, R. Josien, and Y. Choi. 2002. A novel member ofthe leukocyte receptor complex regulates osteoclast differentiation. J. Exp. Med.195: 201–209.

19. Sallusto, F., and A. Lanzavecchia. 1994. Efficient presentation of soluble antigenby cultured human dendritic cells is maintained by granulocyte/macrophagecolony-stimulating factor plus interleukin 4 and downregulated by tumor ne-crosis factor alpha. J. Exp. Med. 179: 1109–1118.

20. Ludewig, B., D. Graf, H. R. Gelderblom, Y. Becker, R. A. Kroczek, and G. Pauli.1995. Spontaneous apoptosis of dendritic cells is efficiently inhibited by TRAP(CD40-ligand) and TNF-a, but strongly enhanced by interleukin-10. Eur. J.Immunol. 25: 1943–1950.

21. Brand, U., I. Bellinghausen, A. H. Enk, H. Jonuleit, D. Becker, J. Knop, andJ. Saloga. 1998. Influence of extracellular matrix proteins on the development ofcultured human dendritic cells. Eur. J. Immunol. 28: 1673–1680.

22. Liu, Z., N. Li, L. A. Diaz, M. Shipley, R. M. Senior, and Z. Werb. 2005. Synergybetween a plasminogen cascade and MMP-9 in autoimmune disease. J. Clin.Invest. 115: 879–887.

23. Pulendran, B., H. Tang, and S. Manicassamy. 2010. Programming dendritic cellsto induce TH2 and tolerogenic responses. Nat. Immunol. 11: 647–655.

24. Hynes, R. O. 2009. The extracellular matrix: not just pretty fibrils. Science 326:1216–1219.

25. Daley, W. P., S. B. Peters, and M. Larsen. 2008. Extracellular matrix dynamics indevelopment and regenerative medicine. J. Cell Sci. 121: 255–264.

26. Karsdal, M. A., T. Woodworth, K. Henriksen, W. P. Maksymowych, H. Genant,P. Vergnaud, C. Christiansen, T. Schubert, P. Qvist, G. Schett, et al. 2011. Bio-chemical markers of ongoing joint damage in rheumatoid arthritis—current andfuture applications, limitations and opportunities. Arthritis Res. Ther. 13: 215.

27. Shortman, K., and S. H. Naik. 2007. Steady-state and inflammatory dendritic-cell development. Nat. Rev. Immunol. 7: 19–30.

28. Leon, B., M. Lopez-Bravo, and C. Ardavın. 2005. Monocyte-derived dendriticcells. Semin. Immunol. 17: 313–318.

29. Steenbakkers, P. G., D. Baeten, E. Rovers, E. M. Veys, A. W. Rijnders, J. Meijerink,F. De Keyser, and A. M. Boots. 2003. Localization of MHC class II/human cartilageglycoprotein-39 complexes in synovia of rheumatoid arthritis patients usingcomplex-specific monoclonal antibodies. J. Immunol. 170: 5719–5727.

30. Merck, E., C. Gaillard, M. Scuiller, P. Scapini, M. A. Cassatella, G. Trinchieri,and E. E. Bates. 2006. Ligation of the FcRg chain-associated human osteoclast-associated receptor enhances the proinflammatory responses of human mono-cytes and neutrophils. J. Immunol. 176: 3149–3156.

31. Baldwin, H. M., T. Ito-Ihara, J. D. Isaacs, and C. M. Hilkens. 2010. Tumournecrosis factor alpha blockade impairs dendritic cell survival and function inrheumatoid arthritis. Ann. Rheum. Dis. 69: 1200–1207.

32. Pai, R. K., D. Askew, W. H. Boom, and C. V. Harding. 2002. Regulation of classII MHC expression in APCs: roles of types I, III, and IV class II transactivator. J.Immunol. 169: 1326–1333.

33. Shin, J. S., M. Ebersold, M. Pypaert, L. Delamarre, A. Hartley, and I. Mellman.2006. Surface expression of MHC class II in dendritic cells is controlled byregulated ubiquitination. Nature 444: 115–118.

34. Chow, A., D. Toomre, W. Garrett, and I. Mellman. 2002. Dendritic cell matu-ration triggers retrograde MHC class II transport from lysosomes to the plasmamembrane. Nature 418: 988–994.

35. Kratz, A., A. Campos-Neto, M. S. Hanson, and N. H. Ruddle. 1996. Chronic inflam-mation caused by lymphotoxin is lymphoid neogenesis. J. Exp. Med. 183: 1461–1472.

36. Page, G., S. Lebecque, and P. Miossec. 2002. Anatomic localization of immatureand mature dendritic cells in an ectopic lymphoid organ: correlation with selectivechemokine expression in rheumatoid synovium. J. Immunol. 168: 5333–5341.

37. Manzo, A., S. Bugatti, R. Caporali, R. Prevo, D. G. Jackson, M. Uguccioni,C. D. Buckley, C. Montecucco, and C. Pitzalis. 2007. CCL21 expression patternof human secondary lymphoid organ stroma is conserved in inflammatorylesions with lymphoid neogenesis. Am. J. Pathol. 171: 1549–1562.

38. Thomas, R., K. P. MacDonald, A. R. Pettit, L. L. Cavanagh, J. Padmanabha, andS. Zehntner. 1999. Dendritic cells and the pathogenesis of rheumatoid arthritis.J. Leukoc. Biol. 66: 286–292.

10 OSCAR-DEPENDENT MATURATION OF DCs

39. Baeten, D., P. G. Steenbakkers, A. M. Rijnders, A. M. Boots, E. M. Veys, andF. De Keyser. 2004. Detection of major histocompatibility complex/humancartilage gp-39 complexes in rheumatoid arthritis synovitis as a specific andindependent histologic marker. Arthritis Rheum. 50: 444–451.

40. Timmer, T. C., B. Baltus, M. Vondenhoff, T. W. Huizinga, P. P. Tak,C. L. Verweij, R. E. Mebius, and T. C. van der Pouw Kraan. 2007. Inflammationand ectopic lymphoid structures in rheumatoid arthritis synovial tissues dissectedby genomics technology: identification of the interleukin-7 signaling pathway intissues with lymphoid neogenesis. Arthritis Rheum. 56: 2492–2502.

41. Krenn, V., N. Schalhorn, A. Greiner, R. Molitoris, A. Konig, F. Gohlke, andH. K. M€uller-Hermelink. 1996. Immunohistochemical analysis of proliferating andantigen-presenting cells in rheumatoid synovial tissue. Rheumatol. Int. 15: 239–247.

42. Li, T. F., J. Mandelin, M. Hukkanen, J. Lassus, J. Sandelin, S. Santavirta,I. Virtanen, and Y. T. Konttinen. 2002. Dendritic cells in rheumatoid synovialmembrane after total removal of the hyaline articular cartilage. Rheumatology(Oxford) 41: 319–323.

43. Sophia Fox, A. J., A. Bedi, and S. A. Rodeo. 2009. The basic science of articularcartilage: structure, composition, and function. Sports Health 1: 461–468.

The Journal of Immunology 11

54

3.2 OSCAR function in monocytes

This part of my studies is being submitted for publication in J. Immunology:

OSCAR-collagen signaling mediates activation of monocytes

Heidi S. Schultz*†1

, Li Guo‡, Pernille Keller

*, Andrew J. Fleetwood

§, Wei Guo

‡, John A. Hamilton

§,

Olle Bjørkdahl*, Martin W. Berchtold

† and Svetlana Panina

*Biopharmaceutical Research Unit, Novo Nordisk A/S, Novo Nordisk Park 1, 2760 Måløv,

Denmark. †Department of Biology, Copenhagen University, Copenhagen 2200, Denmark.

‡Novo

Nordisk Research Centre China, 20 Life Science Park Rd, Changping District, CA 102206 Beijing,

China. §Department of Medicine, University of Melbourne, The Royal Melbourne Hospital,

Parkville, Victoria 3050, Australia.

¶Corresponding author, Svetlana Panina, [email protected], Tel: (+45) 53834930

For supplemental data see Appendix II

1 HSS PhD fellowship was supported by the Danish Agency for Science, Technology, Innovation and Research.

55

3.2.1 Abstract

Osteoclast Associated Receptor (OSCAR) is an activating receptor expressed in human myeloid

cells. Collagen type I (ColI) and collagen type II (ColII) serve as ligands for OSCAR. OSCAR-

collagen interaction stimulates RANK-dependent osteoclastogenesis. We have recently reported

that in monocyte-derived dendritic cells (DC), OSCAR promotes activation and functional

maturation. OSCAR is up-regulated in monocytes from rheumatoid arthritis (RA) patients with

active disease and these monocytes show an increased pro-osteoclastogenic potential. In the current

study, we have addressed a functional role of OSCAR-collagen interaction in monocytes. We show

that the OSCAR-ColII signaling supports survival of monocytes under conditions of growth factor

withdrawal. Moreover, ColII stimulated the release of pro-inflammatory cytokines by monocytes

from healthy donors, which can be completely blocked by an anti-OSCAR monoclonal antibody

(mAb). Mononuclear cells from synovial fluid of RA patients plated on ColII secreted TNF-α and

IL-8 in an OSCAR-dependent manner. Global RNA profiling showed that components of multiple

signaling pathways relevant for the RA pathogenesis are transcriptional targets of OSCAR in

monocytes. Thus, OSCAR plays a pro-inflammatory role in monocyte-derived cells and may

crucially contribute to the RA pathogenesis on multiple levels.

56

3.2.2 Introduction

Rheumatoid arthritis (RA) is a chronic autoimmune disease characterized by a dysregulated

immune response and profound cartilage and bone destruction43

. Infiltration of the synovium by

pro-inflammatory immune cells is the hallmark of RA pathogenesis44;150

. These cells secrete

multiple cytokines and matrix degrading enzymes causing tissue damage. Monocytes are believed

to crucially contribute to the pathophysiology of RA at multiple levels5. The circulating monocytes

of RA patients were shown to have an activated phenotype characterized by production of pro-

inflammatory cytokines IL-1 and IL-6 as well as by up-regulation of the cell surface marker,

integrin CD11b151;152

. Monocytes massively infiltrate the progressively inflamed RA synovium and

produce large amounts of the master pro-inflammatory cytokine, TNF-, thus critically contributing

to the sustained joint inflammation. In the course of inflammation, monocytes serve as precursors

for macrophages, dendritic cells and osteoclasts5. In a number of cross-sectional studies, synovial

sublining monocyte/macrophage infiltration was shown to correlate with scores of disease activity

in RA patients153

. Therapeutic targeting of monocytes/macrophages has been suggested for

treatment of RA154-156

.

The remarkable plasticity of monocytes is influenced by the RA joint

microenvironment, resulting in an altered differentiation, activation and maturation status157

.

Among microenvironmental factors influencing the immune cell activation in RA patients the

extracellular matrix (ECM) proteins has been suggested to be of importance. It has been speculated

that in the joints of RA patients collagens become exposed due to profound synovial architectural

reorganization, thus enabling interaction with ECM receptors137

. Fragments of ColII, the major

collagen of the articular cartilage102

, have been detected in synovial fluid (SF) of patients with joint

diseases158-160

. Several studies have addressed ECM interaction with immune cells. ColI has been

shown to stimulate cytokine secretion by PBMCs100

and by synovial fluid mononuclear cells from

RA patients132

. However, the molecular mechanisms of the ECM-mediated cell activation still

remain obscure. We have recently shown that ColI and ColII stimulate cytokine secretion and

functional maturation of monocyte-derived DCs via the OSCAR signaling pathway142

. OSCAR has

been described as an important player in “osteoimmunology”101

. OSCAR is associated with the

FcRγ136

and functions as a stimulatory co-receptor in osteoclastogenesis signaling via the nuclear

factor of activated T cells c1 (NFATc1) 30

. OSCAR is expressed on all human myeloid cells and is

capable of modulating the immune response11;134;136;142

. The ECM proteins ColI, ColII137;142

and

57

collagen type III137

serve as naturally occurring ligands of OSCAR. It has been proposed that

OSCAR plays a role in skeletal146

and joint disorders139;141

. Monocytes isolated from RA patients

express high levels of OSCAR and show an elevated activation status141

. Furthermore, high levels

of soluble OSCAR, sOSCAR, (presumably, membrane-shed or a secreted splice form) have been

found in plasma, synovial fluid and vasculature of RA patients139-141

. The plasma level of sOSCAR

has been reported to be positively correlating with bone destruction and cardiovascular risk in

patients with RA140

. In contrast, a reverse correlation of sOSCAR present in the serum of RA

patients with the disease activity was reported by the group of Schett141

. Both studies were done on

small patients populations and with a use of different methodologies, thus the prognostic value of

sOSCAR and its function remains unclear.

The function of OSCAR in monocytes has so far been addressed with the use of cross-

linking mAb134

. The authors have shown that triggering of OSCAR induces the release of multiple

chemokines and promotes survival of the cells. Cross-linking of OSCAR in the presence of TLR

ligands, but not alone, potentiated TLR-induced release of pro-inflammatory cytokines134

. In the

current study we addressed the OSCAR-mediated response of monocytes in the presence of the

natural ligand, collagen. We show here that peripheral blood monocytes exposed to ColII increased

their viability in an OSCAR dependent manner. In contrast to the published cross-linking data,

OSCAR engagement by its natural ligand induced profound secretion of pro-inflammatory

cytokines, most of which turned out to be transcriptional targets of the OSCAR-ColII signaling. Our

data suggest that OSCAR-ColII interaction in monocytes may directly contribute to the sustained

inflammation and the joint damage in the RA patients.

58

3.2.3 Materials and methods

Antibodies and ECM proteins

The following monoclonal antibodies (mAb) were used for flow cytometry analysis: Alexa Fluor

647-conjugated anti-OSCAR mAb (Novo Nordisk A/S), APC-conjugated anti-CD14 (BD

Pharmigen), FITC-conjugated anti-CD14 mAb (eBioscience), FITC-conjugated anti-CD68 mAb

(eBioscience), PerCP-Cy5.5–conjugated anti-CD3 mAb (eBioscience), FITC-conjugated anti-HLA-

DR mAb (BD Pharmigen), PE-Cy5-conjugated anti-CD11b mAb (BD Pharmigen), APC-

conjugated anti-TREM1 mAb (BioSite), PE-Cy5-conjugated anti-ICAM1/CD54 mAb (BD

Pharmigen) and PE-conjugated anti-CD86 mAb (IOTest). The isotype control mAb used were anti-

TNP conjugated with Alexa Fluor 647 (Novo Nordisk A/S), IgG1-FITC (eBioscience), IgG2a-

PerCP Cy5.5 (eBioscience), IgG2B-PE (R&D systems), IgG1-APC (BD Bioscience), IgG2b-

FITC (BD Pharmigen), IgG1-PE-Cy5 (BD Bioscience) and IgG1-PE (BD Bioscience).

For functional assays, the OSCAR blocking humanized IgG1.1 mAb (described in Schultz et al.,

2015) was used142

. As an isotype control, an anti-TNP humanized IgG1.1 was used. In some

experiments, an anti-OSCAR non-blocking high affinity humanized IgG1.1 mAb was used as an

isotype control, which produced identical results with the anti-TNP control. In the manuscript, only

anti-TNP control is shown. IgG1.1 antibodies harbor a human Fc fragment with five point

mutations, which prevents activation of the Fc receptors161

.

ColII purified from human cartilage was purchased from Millipore.

Endotoxin level in all protein solutions used for functional studies was measured using the

turbidimetric kinetic LAL assay (Charles River Laboratories). All proteins were essentially

endotoxin-free (below detection limit, <0.01 EU/ml).

Cells

Peripheral blood mononuclear cells (PBMCs) were isolated from leukocyte-enriched buffy coats by

Ficoll-Plague PLUS (GE Healthcare) density centrifugation. Buffy coats from healthy individuals

were obtained anonymously from the Clinical Immunology Blood Bank, The State University

Hospital (Rigshospitalet), Copenhagen, Denmark. All donors gave informed consent according to

the protocol approved by The Ethics Committee for Copenhagen, Denmark, for research use

(ethical approval number H-D-2008-113). Red blood cells (RBC) were removed using a lysis buffer

(eBioscience). Monocytes were purified from PBMCs using CD14 Microbeads human positive

59

selection kit (MACS; Miltenyi Biotec) according to manufactures’ protocol using an AutoMACS

Pro (Miltenyi Biotec). The purity of the CD14+ preparations was assessed by flow cytometry.

Synovial fluid (SF) from RA patients was obtained by needle aspiration anonymously from Peking

University People’s Hospital (PKUPH), Beijing, China. All donors gave informed consent

according to the protocol approved by the Ethics Committee of PKUPH for research use (approval

#20110513-LZ-PKUPH). Synovial fluid cells were isolated by centrifugation at 300xg for 10 min.

Synovial tissue from RA patients was obtained during joint replacement surgery at PKUPH. The

tissue was placed in a buffer (1g tissue per 3 ml buffer) containing 4 mg/ml Collagenase A (Roche)

and 0.1 mg/ml of DNase I (Roche) in RPMI 1640 medium and then homogenized using a MACS

dissociator. The samples were incubated on a MACS mix tube rotator at 37oC for 30 min. The

reaction was stopped by adding an ice-cold growth medium (RPMI 1640 with 10% FCS). The cell

suspension was filtered using a 70μm strainer and centrifuged at 300xg for 10 min. The cells were

re-suspended in a growth medium and processed for flow cytometry analysis.

Activation of cells by plate-bound collagen

Most experiments were performed with both ColI and ColII. Similar to the results obtained with

DCs142

, ColI and ColII had similar effect with ColII being more potent. We therefore show data for

ColII only. For monocytes, 24-well plates were coated with ColII in PBS at ~3µg/cm2

overnight at

4°C. The plate was washed three times in PBS before cells were added to the wells. 1x106

cells

were seeded per well in growth medium (RPMI1640 with GlutaMAX) (Invitrogen, Life

Technologies) supplemented with 10% heat-inactivated FCS and 1% penicillin-streptomycin

(GIBCO, Life Technologies), pre-incubated 30-60 min with anti-OSCAR or isotype control mAb

(1µg/ml) as indicated. The cells were cultured on ColII for various time points as indicated. In some

of the experiments, 100ng/ml LPS (Sigma) and 60ng/ml PMA (Calbiochem) were included as

controls.

RA SF cells (5x105 cells /well) were cultured on high binding 96-well cell culture plates coated

with ColII as described above.

Analysis of cytokine secretion

Supernatants of monocyte cultures were collected after 24 hours of ColII exposure, frozen at -80oC

and shipped to Myriad RBM (Austin, TX) for a 45-Biomarker Multi-Analyte Profiling.

60

Supernatants from 24 hours cultures of SF cells were collected and analyzed using ELISA. The

TNF- ELISA kit was from eBioscience and the IL-8 ELISA kit was from BD Bioscience.

Monocyte viability

Monocytes were washed and re-plated in fresh growth medium without cytokines at 5x105cells/ml.

The monocytes were left untreated or plated onto ColII-coated plates for 2 days in the absence or

presence of anti-OSCAR mAb (1µg/ml) or isotype control (1µg/ml). As a positive control, M-CSF

(10ng/ml) was added. At day two, the cells were washed in PBS and stained with Annexin V-FITC

(BD-Pharmigen) for 15 min at room temperature followed by flow cytometry analysis using a

FACSFortessaTM

(BD Bioscience).

Analysis of cell surface markers by flow cytometry

Expression of cell surface markers was assessed by flow cytometry. In short, cells were washed

twice in PBS with 2% FCS and then stained for 30 min at 37°C in PBS containing 2% FCS in the

presence of 1% human serum albumin (HSA) to block unspecific binding. Cells were rinsed twice

in PBS with 2% FCS prior to flow cytometry analysis. Data acquisition was done using a

FACSFortessaTM

(BD Bioscience). Data analysis was performed using Kaluza software version 1.2

(Beckman coulter).

Microarray analysis

Total RNA was obtained four hours after stimulating the cells with ColII from six donors in the

absence or presence of anti-OSCAR mAb or isotype control mAb. RNA was extracted using TRIzol

(Invitrogen, Life Technologies) followed by purification with an RNeasy MinElute Cleanup Kit

(Qiagen). The RNA integrity was evaluated on an Agilent 2100 Bioanalyser using Agilent RNA

6000 Nano Kit chips (Agilent Technologies), with RNA Integrity Number (RIN) scores of 9.9 or

above. Microarray experiments and data treatment were performed as described in Schultz et al142

.

A paired two-group comparison was done with Benjamini-Hochberg correction for multiple testing

applying a false discovery rate (FDR) of 5% and a biological threshold of more than 2-fold

regulation. The effect of ColII treatment was evaluated using a Principal component analysis

(PCA). Differentially expressed genes were subjected to Gene Ontology (GO) analysis for

biological processes using Ingenuity Pathway Analysis (IPA). The dataset and technical information

61

compliant with MIAME can be found at the EMBL-EBI ArrayExpress depository, accession

number E-MTAB-3281 (http://www.ebi.ac.uk/arrayexpress/experiments/E-MTAB-3281/).

Statistical analysis

Statistical analysis was performed using a Student 2-tailed t tests in a GraphPad Prism 5.0 software.

A P value of less than 0.05 was considered statistically significant. Unless otherwise indicated,

means and standard deviations are shown.

62

3.2.4 Results

OSCAR-collagen interaction supports viability of monocytes

We have previously shown that OSCAR-collagen interaction can promote survival of DC under the

conditions of growth factor withdrawal142

. The group of Merck has shown that cross-linking of

OSCAR promotes survival of monocytes134

. Here, we tested whether the naturally occurring

OSCAR ligand ColII can support monocyte viability. Purified monocytes were cultured on ColII

for two days in the absence of the essential growth factor M-CSF. Cell viability was analyzed by

flow cytometry using Annexin V staining (detecting the cells undergoing apoptosis). Most of the

monocytes deprived of growth factor stained positive for Annexin V on day 2 (Fig.1). All

monocytes cultured in the presence of M-CSF (used as positive control) were Annexin V negative,

i.e. viable. Similarly, ColII seemed to be able to promote survival of the monocytes, as

approximately 80% of the cells stained negative for Annexin V. When OSCAR-ColII signaling was

blocked with an anti-OSCAR mAb cell viability was drastically decreased, which was not the case

when an isotype control mAb was used.

Fig. 1. ColII supports survival of monocytes at the conditions of growth factor withdrawal. Freshly purified

monocytes were cultured on either uncoated or ColII-coated plates. To the ColII coated wells, OSCAR blocking mAb

(1µg/ml) or control mAb (1µg/ml) were added as indicated. M-CSF (10ng/ml) was used as positive control. As a

negative control monocytes were left in growth factor free medium. After 2 days, cell viability was assessed using

Annexin V staining. For flow cytometry analysis, single cells were gated in a SSC-H vs SSC-W scatter (Gate A), then

debris was gated out in a SSC vs FSC scatter (Gate B). Histograms are shown for the cells analyzed within gate B.

Numbers in the corners show the percentage of cells negative for Annexin V (viable). Data are shown for one

representative donor out of four analyzed.

OSCAR-collagen signaling stimulates secretion of pro-inflammatory mediators

We investigated the effect of OSCAR-ColII interaction on the ability of monocytes to secrete pro-

inflammatory mediators. Freshly purified monocytes from six healthy donors were stimulated with

63

ColII for 20 hours and the cell culture supernatants were analyzed for the presence of inflammation

markers. An array of 45 different factors, such as cytokines, chemokines and acute-phase reactants,

was analyzed in the cell culture supernatants using the Multi-Analyte Profiling (MAP) technology

platform (service provided by Myriad RBM Inc.). In the supernatants, 19 inflammation markers

were detected, of which 14 were significantly up-regulated upon ColII exposure as compared to the

untreated control and five proteins showed a tendency to be up-regulated. Stimulation with ColII

triggered production of Alpha-1-Antitrypsin (AAT), Beta-2-Microglobulin (B2M), Fibrinogen, IL-

1, IL-1ra, IL-8, MIP-1, MIP-1, Vascular endothelial growth factor (VEGF), RANTES, tissue

inhibitor or metalloproteinases (TIMP)-1, TNF-, TNFR2 and vitamin D binding protein (VDBP)

in monocytes (Table I), which were down-regulated to the baseline level when OSCAR-collagen

interaction was blocked by an anti-OSCAR mAb, but not by the isotype control. The following

markers, ICAM1, IL-6, IL-10, IL-18 and MCP-1, showed a tendency to be up-regulated when

stimulated with ColII, but the data did not reach a statistical significance.

Table I. Inflammatory mediators secreted by monocytes exposed to ColII

monocytes ColII-monocytes

ColII-monocytes + anti-

OSCAR

ColII-monocytes +

isotype

AAT (ng/ml) 79.0±4.8 92.0±4.9* 67.0±3.0** 96.0±7.1*

B2M (µg/ml) 0.12±0.013 0.16±0.014** 0.10±0.011* 0.16±0.017**

Fribrinogen (ng/ml) 6.9±0.94 10±2.1* 7.2±1.0 10±2.3

ICAM-1 (ng/ml) 0.49±0.018 0.67±0.10 0.53±0.027 0.59±0.048*

IL-1b (pg/ml) 6.3±2.7 15±5.3* 2.9±0.55 14±4.4**

IL-1ra (pg/ml) 147±32 797±213* 90±14* 807±198*

IL-6 (pg/ml) 19±11 51±24 17±6.3 41±18

IL-8 (ng/ml) 0.39±0.18 1.4±0.57* 0.21±0.048 1.2±0.39*

IL-10 (pg/ml) 3.4±0.82 4.8±1.2 3.3±0.58 4.9±1.3

IL-18 (pg/ml) 6.3±1.2 7.8±1.3 5.4±1.0 7.1±1.3

MIP-1a (pg/ml) 24±8.3 58±11*** 14±3.7 60±9.6***

MIP-1b (pg/ml)) 153±50 331±52** 108±28 334±54***

MCP-1 (pg/ml) 17±5.9 29±9.1 15±3.2 30±8.4

RANTES (pg/ml) 3.2±0.39 4.2±0.59* 2.6±0.32*** 3.9±0.42**

TIMP-1 (ng/ml) 22±6.0 54±14* 13±2.5 55±15*

TNF-a ((pg/ml) 22±7.4 89±23** 11±2.0 85±19**

TNFR2 (ng/ml) 2.1±0.23 3.2±0.26*** 1.4±0.13** 3.2±0.2**

VEGF (pg/ml) 307±83 419±83** 212±31 367±61

VDBP (ng/ml) 1.2±0.13 1.4±0.16* 1.1±0.11 1.4±0.14*

Ferritin (ng/ml) 101±16 84±11 92±16* 88±10

IL-1a (pg/ml) 0.63±0.11 0.66±0.11 0.43±0.036 0.7±0.083

CD14+ monocytes were either pre-incubated with 1µg/ml anti-OSCAR mAb or an isotype control for 30 min before plating onto

ColII-coated plates for 20 hours. Supernatant was collected, frozen and shipped to Myriad RBM in Austin, TX for a 45-biomarker

Multi-Analyte profiling. AAT: Alpha-1-Antitrypsin; B2M: Beta-2-Microglobulin; MIP, Macrophage Inflammatory Protein; TNFR2:

Tumor Necrosis Factor Receptor 2; TIMP-1: Tissue Inhibitor of Metalloproteinases 1; VEGF, Vascular Endothelial Growth Factor;

VDBP: Vitamin D-Binding Protein. p<0.05=*; p<0.01=**; p<0.001=*** indicate significant difference from untreated.

64

OSCAR expressing cells are present in the RA joints and respond to stimulation by collagen

The expression of OSCAR was analyzed in the SF mononuclear cells as well as in the cells from ST

of RA patients by flow cytometry. OSCAR was expressed on myeloid cells present in SF (CD14+

and CD86+ cells) but not on lymphocytes (CD3

+ cells) of RA patients (Fig.2A), which is consistent

with the previously published data on peripheral blood cells from healthy donors136

. In ST of RA

patients, OSCAR expression was detected on CD14+, CD68

+and CD86

+ cells (Fig.2B). Culturing

SF mononuclear cells from RA patients on ColI has previously been shown to induce secretion of

IL-1, IL-6 and TNF-132

, however, the molecular mechanisms controlling such cell activation

have not been described so far. Here, RA synovial fluid cells were cultured on plate-bound ColII in

the presence or absence of an anti-OSCAR antagonistic mAb or an isotype mAb. After one day, cell

culture supernatants were collected and cytokines were analyzed by ELISA. The natural ligand of

OSCAR, ColII, induced TNF- and IL-8 cytokine secretion from RA SF cells, which was

significantly inhibited by the anti-OSCAR mAb, but not by the isotype control mAb (Fig.3). Our

data suggest that OSCAR-expressing myeloid cells are present in the joints of RA patients and can

potentially be activated by collagen via the OSCAR signaling pathway.

Fig. 2. OSCAR-expressing cells are present in synovial fluid (A) and synovial tissue (B) of RA patients. For flow

cytometry analysis, cell debris was gated out using FSC vs. SSC; the gate was then set for single cells, which were

further analyzed using multicolor staining for the indicated markers. OSCAR expression was analyzed in cell

populations gated for CD14+, CD86+ or CD3+ (for SF cells) or on the cell populations gated for CD14+, CD86+ and

CD68+ in case of ST. Data for one representative patient out of 10 analyzed are shown.

65

Untr

eate

dColII

ColII

+ 1

0 µg

/ml a

-OSCAR

ColII

+ 1

0 µg

/ml I

soty

pe

0

200

400

600

800

1000A

P=0.0001

P=0.0001

TN

F-

(p

g/m

l)

Untr

eate

dColII

ColII

+ 1

0 µg

/ml a

-OSCAR

ColII

+ 1

0 µg

/ml I

soty

pe

0

3

6

9

12B

P=0.0001

P=0.0001

IL-8

(n

g/m

l)

Fig. 3. ColII induces inflammatory cytokine secretion from RA synovial fluid cells. Synovial fluid cells were plated

on ColII for 20-24 hours. The supernatants were analyzed for TNF-α (A) and IL-8 (B) using ELISA. Data are expressed

as mean ± SD and shown for 1 representative patient out of 5 being analyzed. .

OSCAR engagement by collagen induces up-regulation of activation markers expressed at the

cell surface

We have further analyzed ColII- exposed monocytes derived from healthy donors for the expression

of the surface markers, including the activation marker integrin CD11b151;152

, the pro-inflammatory

marker Triggering Receptor Expressed on Myeloid cells (TREM-1)162

, and the immune synapse

components HLA-DR, ICAM-1 and CD8617

. The monocytes were cultured on ColII for four hours

before the expression of surface markers was examined by flow cytometry. We observed that ColII

was able to induce up-regulation of the cell surface markers TREM-1, CD11b, ICAM-1 and CD86,

but not HLA-DR (Fig.4). The up-regulation of these surface markers by ColII was abolished when

OSCAR-ColII interaction was blocked with an anti-OSCAR mAb, whereas the isotype control mAb

had no effect.

66

Untr

eate

dColII

ColII

+ 1

µg/m

l a-O

SCAR

ColII

+ 1

µg/m

l iso

type

LPSPM

A

0

10

20

30

40

50

p=0.1

HL

A-D

R (

MF

I)

Untr

eate

dColII

ColII

+ 1

µg/m

l a-O

SCAR

ColII

+ 1

µg/m

l iso

type

LPSPM

A

0

10

20

30

p=0.01

p=0.002

p=0.01

TR

EM

1 (

MF

I)

Untr

eate

dColII

ColII

+ 1

µg/m

l a-O

SCAR

ColII

+ 1

µg/m

l iso

type

LPSPM

A

0

50

100

150

200

250

p=0.02

p=0.04

p=0.01

p=0.02

CD

11

b (

MF

I)

Untr

eate

dColII

ColII

+ 1

µg/m

l a-O

SCAR

ColII

+ 1

µg/m

l iso

type

LPSPM

A

0

5

10

15

20

p=0.01

p=0.003

p=0.004

CD

86

(M

FI)

Untr

eate

dColII

ColII

+ 1

µg/m

l a-O

SCAR

ColII

+ 1

µg/m

l iso

type

LPSPM

A

0

50

100

150

p=0.01

p=0.01

p=0.002

p=0.002

ICA

M1

(M

FI)

Fig. 4. Cell surface markers expressed on the ColII-cultured monocytes. 4x106 monocytes (CD14

+) were pre-

treated with either 1µg/ml anti-OSCAR mAb or 1µg/ml isotype control mAb for 30 min, transferred onto ColII-coated

plates and incubated for four hours. As positive controls, LPS (100ng/ml) and PMA (60ng/ml) were used. Cells were

analyzed by flow cytometry for surface expression of HLA-DR (FITC), TREM (APC), CD11b (PE-Cy5), ICAM1 (PE-

Cy5) and CD86 (PE). For flow cytometry analysis, single cells were gated in a SSC-H vs SSC-W scatter, followed by

gating out cell debris in a SSC vs FSC scatter. The expression level of the respective surface markers is presented as

MFI values (median fluorescence intensity). Data analysis was performed using a Kaluza software. Data represent mean

±SD for four donors analyzed.

Signaling by OSCAR-collagen regulates gene expression in monocytes

To elucidate the molecular mechanisms of the OSCAR-Col effects on monocytes, a gene expression

profiling was performed using a full transcriptome Affimetrix gene chip microarray. Monocytes

were cultured for 4 hours on uncoated plates or on ColII-coated plates in the presence or absence of

anti-OSCAR mAb or isotype control mAb. The full microarray dataset can be downloaded from the

ArrayExpress depositary (accession number E-MTAB-3281). ColII treatment led to significant

67

changes in the expression level of 226 genes (>2 fold change; 5% false discovery rate;

Supplemental Table SIA). Treatment of the ColII exposed cells with an anti-OSCAR mAb had an

effect on the expression of 528 genes when compared to the treatment with an isotype mAb

(Supplemental Table SIB). Of the 226 genes regulated by ColII, 218 were clearly regulated via the

OSCAR signaling pathway as the ColII effect was fully abolished by the treatment with an anti-

OSCAR mAb and not by the isotype mAb (“true OSCAR-collagen target genes”; Supplemental

Table S1C). The OSCAR-ColII target genes were clustered into functional categories and pathways

using the Ingenuity Pathway Analysis. The top 20 canonical pathways are listed in Table II. We

found that the OSCAR-ColII signaling regulates the expression of genes involved in

adhesion/diapedesis, cell activation, immune regulation and a number of genes implicated in

various signal transduction pathways implicated in inflammation. OSCAR seems to be involved in

the transcriptional regulation of pro-inflammatory cytokines, their receptors and components of

their signaling pathways. Among those, expression of the gene for IL-1 (IL1A), its receptor IL1R1

as well as the receptor antagonist IL1RN (coding for the IL-1Ra protein) was up-regulated.

OSCAR-ColII up-regulated expression of the gene coding for the key pro-inflammatory cytokine

TNF-α as well as the TNF family member TNFSF14 (coding the LIGHT protein), while expression

of another TNF-related gene, TNFSF10 (coding TRAIL) was down-regulated. A number of genes

encoding chemokines (CCL7, CXCL1, -2 and -3, CXCL5, CCL20 and CCL24) were found to be

transcriptional targets of the OSCAR-ColII signaling. The most pronounced up-regulation was seen

for the gene encoding CCL24, a chemoattratctant for eosinophils, neutrophils and macrophages 163

.

Other OSCAR-ColII regulated molecules and pathways with relevance to inflammation included

components of the NF-κB, TREM-1, Jak2/STAT and TLR signaling pathways as well as ECM

modifying enzymes such as MMP7 and MMP19. A number of genes were down-regulated by the

OSCAR-ColII signaling, including FZD2 (encoding frizzled 2, a receptor for WNT proteins) as

well as chemokine receptors, CCR2 and CX3CR1.

68

Table II. Top 20 canonical pathways regulated by the OSCAR-ColII signaling

Ingenuity Canonical Pathways Molecules -Log (p-value)

Granulocyte Adhesion and Diapedesis MMP7,IL1A,CCL20,CCL24,IL1R1,CXCL5,SDC4,CXCL3,I

L1RN,SDC2,CXCL1,CXCL2,TNF,MMP19,CCL7

9,72E+00

Agranulocyte Adhesion and Diapedesis MMP7,IL1A,CCL20,CCL24,IL1R1,CXCL5,SDC4,CXCL3,I

L1RN,CXCL1,CXCL2,TNF,MMP19,CCL7

8,38E+00

Atherosclerosis Signaling IL1A,IL1RN,LPL,APOC1,CCR2,F3,TNF,TNFSF14,TNFRS

F12A

5,52E+00

Role of IL-17A in Psoriasis CXCL3,CCL20,CXCL1,CXCL5 5,27E+00

TREM1 Signaling CXCL3,NLRP12,TLR8,TLR7,JAK2,TNF,CCL7 5,09E+00

LXR/RXR Activation IL1A,IL1RN,LPL,SERPINF1,APOC1,IL1R1,TNF,CCL7 4,50E+00

Role of Macrophages, Fibroblasts and

Endothelial Cells in Rheumatoid Arthritis

IL1A,F2RL1,RRAS,IL1RN,TLR8,TLR7,CEBPD,IL1R1,JA

K2,TNF,FZD2

3,77E+00

Role of IL-17A in Arthritis CXCL3,CCL20,CXCL1,CXCL5,CCL7 3,71E+00

Altered T Cell and B Cell Signaling in

Rheumatoid Arthritis

IL1A,SPP1,IL1RN,TLR8,TLR7,TNF 3,70E+00

Acute Phase Response Signaling IL1A,RRAS,IL1RN,SERPINF1,IL1R1,JAK2,SERPINE1,TN

F

3,62E+00

NF-B Signaling IL1A,RRAS,IL1RN,TLR8,TLR7,MAP3K8,IL1R1,TNF 3,58E+00

PPAR Signaling PPARG,IL1A,RRAS,IL1RN,IL1R1,TNF 3,55E+00

FXR/RXR Activation PPARG,IL1A,IL1RN,LPL,SERPINF1,APOC1,TNF 3,44E+00

IL-17A Signaling in Airway Cells CXCL3,CCL20,CXCL1,CXCL5,JAK2 3,38E+00

Colorectal Cancer Metastasis Signaling MMP7,RRAS,TLR8,TLR7,RHOU,JAK2,TNF,FZD2,MMP1

9

3,28E+00

Cholecystokinin/Gastrin-mediated

Signaling

IL1A,RRAS,IL1RN,CREM,RHOU,TNF 3,27E+00

Toll-like Receptor Signaling IL1A,IL1RN,TLR8,TLR7,TNF 3,15E+00

Role of Tissue Factor in Cancer F2RL1,RRAS,CXCL1,JAK2,F3,EIF4E 3,14E+00

IL-6 Signaling IL1A,RRAS,IL1RN,IL1R1,JAK2,TNF 3,06E+00

HMGB1 Signaling IL1A,RRAS,RHOU,IL1R1,SERPINE1,TNF 2,99E+00

Monocytes were cultured on ColII for 4 hours in the presence or absence of anti-OSCAR mAb (1µg/ml) or isotype control

(1µg/ml). Total cellular RNA was isolated and genome-wide mRNA expression analysis was done using Affymetrix GeneChip

microarray. The table shows canonical pathways for the OSCAR-ColII regulated genes at a false discovery rate at 5% and a

biological criterion of minimum two-fold change. A complete list of all genes regulated by OSCAR-ColII signaling is provided in

Supplemental Table SIC. Samples from six independent donors were analyzed.

69

3.2.5 Discussion

Differentiation, polarization and activation of leukocytes have been shown to be influenced by their

interaction with ECM proteins10;100;164

, however it still remains unclear which receptors mediate

these effects. We have recently shown that ColI and ColII induce functional maturation and

activation of human monocyte-derived DCs through the OSCAR signaling142

. Besides providing

structural support, collagens exert an important role in various cellular functions, including

proliferation, migration and apoptosis165

. The structure of ECM is pivotal for defining the cellular

response166

. In RA, joint destruction is a hallmark of the disease. The structural damage leads to an

increased turnover of collagens within the joint167

. This makes collagens accessible for interaction

with the infiltrating immune cells, which can alter the cell differentiation and activation status.

Monocytes, the major drivers of synovial inflammation, infiltrate the RA synovial tissue where they

encounter collagen in their environment. OSCAR-expressing mononuclear cells have been detected

in the inflamed synovium in contact with the exposed collagens137;141

. This let us hypothesize that

the activation status of monocytes exiting the circulation might be modulated by the exposed

collagens in the RA joint microenvironment via OSCAR signaling.

Merck and colleagues have shown that cross-linking of OSCAR promotes survival of

monocytes134

. Here, we examined the functional role of OSCAR stimulation with ColII in

monocytes. We show that peripheral blood monocytes exposed to ColII exhibited increased their

viability in the absence of growth factors, and that this effect was mediated by the OSCAR

pathway. Though the molecular mechanisms of this effect were not addressed in the current

manuscript, we speculate based on the microarray data that the OSCAR-collagen signaling

regulates the expression of genes involved in the control of cell death and survival. Among

transcriptional targets of OSCAR are genes encoding components of the NFB- and ERK/MAPK

signaling machinery (both can work in pro-survival pathways), as well as members of the death

receptor signaling like TNF-Related Apoptosis Inducing Ligand (TNFSF10 or TRAIL) and TNF-.

TRAIL triggers programmed cell death and is down-regulated in monocytes stimulated with

collagen. TRAIL has been linked to the pathogenesis of RA where it has been suggested to inhibit

the development of the disease by affecting the viability of the cells in the synovium168

. TNF- is

known to induce survival of myeloid cells169-171

. We observed an up-regulation of TNF- at both

gene and protein levels, possibly contributing to the increased viability of monocytes in an

autocrine and paracrine manner.

70

We detected secretion of several inflammatory markers when monocytes were

exposed to ColII, indicating that collagen activates monocytes in an OSCAR-dependent manner. In

contrast to the data published by Merck et al. reporting that OSCAR cross-linking cannot induce

secretion of pro-inflammatory cytokines, but chemokines134

, we detected a profound secretion of

several pro-inflammatory cytokines including IL-1β, IL-6 and TNF-α. These cytokines are all the

key players in RA pathogenesis and have been found in high concentrations in the synovial fluid of

RA patients25

. IL-1, IL-6 and TNF- are able to increase cellular infiltration of the synovium,

potentiate osteoclastogenesis and induce MMP production within the RA joint25;110;172

. TNF- is

also seen to enhance OSCAR expression on monocytes141

, which might amplify the effect via a

positive feedback mechanism.

Several chemokines were induced by the OSCAR-ColII interaction. Among the

secreted chemokines, IL-8 (recruiting granulocytes), RANTES (recruiting T cells) and MIP-

1α/βchemoattractant for monocytes/macrophages, NK cells and neutrophils) were significantly

up-regulated. Moreover, MIP-1/ has been shown to stimulate the production of reactive oxygen

species and to induce the synthesis of pro-inflammatory cytokines in macrophages and

fibroblasts173-175

. At the transcriptional level, several chemokine genes were up-regulated including

CCL24 (recruiting resting T cells and, to a lesser degree, granulocytes), CXCL3 and CXCL5 (both

attracting neutrophils and acting on the same receptor, CXCR2) and CCL7 (recruiting

monocytes/macrophages). Altogether, these data indicate the potential of OSCAR signaling to

contribute to the amplification of the inflammatory response in RA.

The monocyte chemokine receptors, CCR2 (receptor for MCP1) and CX3CR1

(fractalkine receptor) were down-regulated at the transcriptional level upon exposure of the cells to

ColII in an OSCAR-dependent manner. These receptors, which are critical for the initial

recruitment of monocytes, have been reported to be down-regulated upon extravasation of

monocytes into the inflamed tissue176;177

and their sequential differentiation to macrophages178

,

which might be essential for cell retention within the inflamed tissue. It is likely OSCAR-ColII

signaling is involved, at least in part, in these previously described phenomena.

The expression of several cell surface markers including CD11b, TREM-1, ICAM1

and CD86 was found to be up-regulated in ColII-exposed monocytes in an OSCAR-dependent

manner. CD11b has been shown to be an activation marker in monocytes, which expression was

enhanced in RA patients. It has been suggested that CD11b is involved in monocyte infiltration of

the synovium151

. We have previously shown that OSCAR-collagen signaling is controlling the

71

expression of the major immune synapse components (HLA-DR, CD40, CD80, CD86 and ICAM1)

in monocyte-derived DCs142

. Similar to DCs, monocytes may act as APC, though being less potent.

Here, we show that, unlike DCs, ColII-exposed monocytes did not up-regulate expression of HLA-

DR, though the expression of the adhesion molecule ICAM1 and the co-stimulatory molecule CD86

was increased upon OSCAR engagement by ColII. Thus, OSCAR might potentially contribute to

the monocyte-T cell interaction via stabilization of cell-cell interaction and by enhancing co-

stimulatory signaling.

Another surface marker that was induced by the OSCAR-ColII interaction on the

transcript and protein expression level was TREM-1. The Ingenuity Pathway Analysis (Table II)

has revealed that multiple downstream components of the TREM-1 pathway were up-regulated by

the OSCAR-ColII signaling in monocytes. TREM-1 is a member of the immunoglobulin protein

superfamily that activates myeloid cells through the DAP12 signaling. TREM-1 has been shown to

amplify TLR and NOD-mediated signaling, where it synergistically enhances cytokine

secretion162;179-182

. TREM-1 expression is controlled by the PI3K- and NFB pathways180;183

which

seem to be regulated by the OSCAR signaling. TREM-1 has mainly been implicated in bacterial

infections180

. Besides, TREM-1 expression was also shown to be up-regulated in the RA

synovium184

and blocking TREM-1 signaling in CIA mice models significantly improved the

disease outcome185

indicating that TREM-1 is associated with the RA pathogenesis. We suggest that

the OSCAR pro-inflammatory effect can, at least in part, be mediated by activation of the TREM-1

signaling cascade (Supplemental Figure S1).

We have identified an OSCAR-driven up-regulation of the genes encoding MMP7 and MMP19 in

monocytes cultured on ColII. MMP19 has been detected at the surface of activated peripheral blood

mononuclear cells186

and in blood vessels of the inflamed synovium of RA patients187

. MMP-19 is

able to hydrolyze basement membrane components188

and constituents of cartilage189

. This suggests

that OSCAR-ColII interaction may potentially enhance extravasation and diapedesis of monocytes

during arthritic diseases, leading to an increased invasion of the pannus into the joint, thus

contributing to joint destruction.

In RA, the level of OSCAR expression on monocytes has been shown to positively

correlate with the disease activity and it was suggested that OSCAR functions as a co-stimulatory

receptor enhancing osteoclastogenesis from monocyte-derived precursors 141

. In the synovial tissue

of RA patients, OSCAR expressing mononuclear cells were found to be localized adjacent to the

72

microvasculature139;141

. Here, we tested whether cells present in the synovial fluid of RA patients

can be activated by collagen via OSCAR. The mononuclear SF cells stimulated with ColII secreted

TNF- and IL-8, which was abrogated by blocking the OSCAR-ColII interaction with an anti-

OSCAR antagonistic mAb. Our results support previously published findings on the activation of

mononuclear cells from RA patients cultured on collagen132

and identify OSCAR as being the

receptor mediating this effect. Thus, OSCAR is functional in mononuclear cells present in the SF of

RA patients and these cells - via the OSCAR-ColII interaction - can contribute to the RA

pathogenesis by secreting pro-inflammatory cytokines and chemokines. Collagen fragments have

been identified in the synovial fluid of RA patients158;159;190;191

. Though we did not test whether they

can serve as functional OSCAR ligands, we speculate that they may potentially contribute to

activation of the myeloid cells present in the synovial fluid.

The present study further elucidates the role of OSCAR in myeloid cells. Similar to its

function in DCs142

, the primary function of OSCAR in monocytes seems to be pro-inflammatory.

Our data suggest that the OSCAR-ColII interaction in monocytes may be important in perpetuating

the course of the disease in RA patients. Targeting of OSCAR-ColII signaling may thus represent a

novel approach for therapy of RA.

73

3.3 OSCAR function in macrophages

3.3.1 Introduction

Macrophages are one of the most abundant cell types in the inflamed synovial membrane. The

number of macrophages correlates with damage, pain and inflammation of the joint, indicating an

important role in RA pathogenesis154;192;193

. Macrophages are essential for homeostasis of the body

where they scavenge debris made by physiological or pathological processes. Activation of

macrophages has been observed in RA tissue6. When activated, macrophages secrete a broad

repertoire of proteases, reactive oxygen species, pro-inflammatory cytokines and chemokines. That

in turn recruit and activate other immune cells, thus contributing considerably to the sustained

inflammation and joint destruction in RA6;194

. One of the characteristic features of RA is bone and

cartilage destruction. Macrophages can contribute to the joint destruction by releasing cytokines

such as IL-1 and TNF-, and matrix metalloproteinases (MMPs) or by activating fibroblasts to

produce MMPs43

. Macrophages can differentiate into bone degrading osteoclasts within the RA

synovium under the influence of factors like RANKL and M-CSF26

.

OSCAR is expressed on all myeloid cells in humans, including macrophages135;136

. In

DCs and monocytes OSCAR has been described to have an immunomodulating

role11;134;136;137;141;142

. However, no study has described the function of OSCAR in macrophages. In

house data indicated that cross-linking OSCAR on macrophages induces secretion of MMP-1, -3, -

7, -8 and 12 (preliminary data by S. Panina, Novo Nordisk A/S). In DCs, we have shown by

microarray analysis that MMP-9, MMP-10 and the uPA receptor (uPAR) are transcriptional targets

of the OSCAR signaling142

. Therefore, in this study, we investigated the hypothesis that OSCAR

might be involved in the control of ECM degradation by macrophages and that blocking of OSCAR

signaling could hinder joint damage.

74

3.3.2 Material and methods

Antibodies and ECM proteins

For functional assays several OSCAR-blocking mAb were generated by recombinant technology at

Novo Nordisk A/S (described in Schultz et al., 2015)142

. As an isotype control, anti-TNP human

IgG1.1 was used. IgG1.1 antibodies harbor a Fc fragment with five point mutations that abolish

binding to Fc receptors, as described in161;195

.

ColII purified from human cartilage was purchased from Millipore. Endotoxin level in the protein

solutions was measured using the turbidimetric kinetic LAL assay (reagents supplied by Charles

River Laboratories). All proteins used for functional assays in this study were essentially endotoxin-

free (below detection limit of 0.01 EU/ml).

Macrophage generation and culturing on collagen

Buffy coats from healthy individuals were obtained anonymously from Rigshospitalet Blood Bank,

Copenhagen. All donors gave informed consent according to the protocol approved by The Ethics

Committee for Copenhagen, Denmark, for research use (approval #H-D-2008-113). PBMCs were

purified by Ficoll-Plaque density centrifugation. Monocytes were purified from PBMCs using the

CD14+ Microbeads positive selection kit (Miltenyi Biotec). To induce macrophage differentiation,

purified monocytes were seeded at 3x105 cells/ml in RPMI1640 (Gibco, Life technologies)

supplemented with 10% (v/v) fetal bovine serum, Glutamax (Gibco, Life technologies), 1%

penicillin/streptomycin (Gibco, Life technologies) (referred to as medium), and recombinant human

macrophage colony stimulating factor (CSF-1; Chiron, 2000U/ml or M-CSF, RD systems, 40ng/ml)

(Science Products Dept. or Nunclon) for six days in a 5% CO2 tissue culture incubator at 37°C. At

day three, the medium was replenished. To loosen macrophages from the dish the media was

removed, the cells were quickly flushed with 10ml ice-cold PBS, before adding 10ml PBS

containing 5mM EDTA to the macrophages. The cells were left for 30 min in a 5% CO2 tissue

culture incubator at 37°C. 1x106

cells were thoroughly washed in medium followed by a 20 min

pre-incubation with 1µg/ml OSCAR-blocking, control mAb or “vehicle” (medium without Ab).

Untreated virgin polysterene 24-well plates (Science Products Dept.) were coated with ColII in PBS

at (~3µg/cm2) overnight at 4°C. The plate was washed three times in PBS before cells were added to

the wells. The cells were cultured on ColII for 4 hours or 20 hours as indicated.

75

Analysis of cell surface markers by flow cytometry

Expression of cell surface markers was assessed by flow cytometry. Cells were washed twice in

PBS with 2% FCS and stained with anti-OSCAR-Phycoerythrin (PE)-conjugated mouse mAb

(clone 11.1CN5, Beckman Coulter) or murine IgG2B-PE isotype control (R&D systems) for 30 min

at 37°C in PBS with 2% FCS in the presence of 1% HSA to block unspecific binding. Cells were

washed three times in PBS with 2% FCS prior to flow cytometry analysis. Data acquisition was

performed using a FACSFortessaTM

(BD Bioscience). APC-H7 Dead Cell Stain (Invitrogen, Life

technologies) was used for dead cell exclusion. For compensation, individual Ab stain was used

with one drop of negative control beads and anti-mouse IgG beads (BD Biosciences). For the dead

cell marker compensation arC reactive beads were used (Invitrogen, Life technologies). Data

analysis was performed using Kaluza software version 1.2 (Beckman coulter).

Degradation assay

Macrophage matrix degradation was analyzed using a QCMTM

Gelatin Invadopodia assay (Merck-

Millipore) according to the manufacturer’s protocol. Briefly, poly-L-lysine was absorbed to the

glass coverslip and treated with glutaraldehyde before a layer of FITC-conjugated gelatin was

added to the surface. Subsequently, the fluorescently-coated glass was disinfected with 70% ethanol

and left with medium for 30 min before seeding the cells. Macrophages (3x104 cells) were cultured

on the gelatin surface in the presence of M-CSF (1000U/ml), for 24 hours in a 5% CO2 tissue

culture incubator at 37°C. Following incubation, the cells were fixed with 3.7% formaldehyde and

stained for F-actin (TRICT-Phalloidin) and nuclei (DAPI). Prior to imaging, the coverslips were

transferred to glass slides, containing 1 drop of mounting gel (ProSciTec), sealed and left in dark at

4oC to settle overnight. Cells were imaged using a fluorescence microscope (Zeiss Axioskop 2; Carl

Zeiss; at original magnificationx20). Images were captured by a Zeiss AxioCam MRm (Carl Zeiss)

and visualized with AxioVision 4.4 software. Quantification of the degradation area in percentages

was performed in ImageJ, version 1.46r (NIH, Rasband). The Images were changes to 8 bit images

and the threshold was defined specifically for each donor. The percentage of degradation was

normalized to number of untreated cells. In some experiments, matrix degradation was measured in

the presence of 250nM Plg (Enzyme Research Lab), 10nM aprotinin (Sigma-Aldrich), and 1µg/ml

anti-OSCAR mAb or isotype control.

76

Note, that the manufacturer has disclosed that gelatin used in the assay was extracted from porcine

skin using an acid based method and is known as Type A. Though, they denied disclosing the

composition of the gelatin.

Assessment of MMP secretion

Human MMP base kit Luminex® Performance Assay was used to analyze the quantity of selected

MMPs in supernatants from macrophages, according to manufacturer’s recommendations. In short,

MMP Standard cocktail was reconstituted and diluted in 500μl of Calibrator Diluent RD5-37. After

15 minutes a 3x serial dilution was performed in Calibrator Diluent RD5-37. Standards were ran in

duplicates. 50µl of the microparticle mixture was added to the pre-wetted filter-bottomed

microplate, followed by addition of 50µl of sample or standard. It was left to incubate for 2 hours at

RT on a shaker (500rpm) protected from light. Using a vacuum manifold device the plate was

washed three times in the supplied wash buffer (100µl). Then 50µl Biotin Antibody Cocktail was

added and incubated at the same conditions for one hour. The wash step was repeated. Diluted

streptavidin was added to the wells and incubated at the same conditions for 30min. 100μl Wash

buffer was added per well and incubated on the shaker for 2 min before reading the plate using the

BioPlexTM

200 System (Luminex). Data were analyzed using GraphPad Prism 5.0.

Assessment of cytokine secretion

Multiplex analysis of cytokines and chemokines in the cell culture supernatants was performed

using the BioPlexTM

200 Cytokine Assay platform (Bio-Rad), and a Luminex 200 instrument

(Luminex), according to manufacturer’s protocols. Data was analyzed using GraphPad Prism 5.0.

MMP activity

To measure MMP activity a fluorogenic substrate for several matrix metalloproteinases (MMPs)

(Enzo Life Sciences) was used according to manufacturer’s protocol. Briefly, the peptide was first

dissolved in DMSO to a concentration of 20mM and then diluted in assay buffer (500mM HEPES,

100mM CaCl2 and 0.5% Brij-35 – pH 7) to a 10µM concentration. 100µl of the substrate was added

to a Black microwell 96 well (NUNC) together with 100µl of sample and left to incubate for 1h at

37oC. The Fluorescence intensity was measured using a Paradigm microplate reader (Beckman

Coulter) at the excitation/emission values of 360/450nm set to 37oC every 10 minutes for 10 cycles.

77

Microarray analysis

Total RNA was obtained four hours after culturing macrophages on ColII in the absence or

presence of anti-OSCAR mAb or isotype control mAb. Cells were lysed in 400l TRIzol

(Invitrogen, Life technologies). After adding 80µl chloroform, samples were shaken vigorously for

15 seconds and incubated at room temperature for 3 minutes. The lysate was then centrifuged at

12,000xg for 15 min at 4oC and RNA was purified from the aqueous phase using a RNeasy

MinElute Cleanup Kit (Qiagen, Hilden) according to the manufacturers’ protocol. Subsequent,

RNA was eluted in 14l water. RNA integrity was evaluated on an Agilent 2100 Bioanalyser using

Agilent RNA 6000 Nano Kit chips (Agilent Technologies); RNA Integrity Number (RIN) scores

were of 9.9 or above. The RNA concentration was measured using a NanoDrop® ND-1000

(Saveen&Werner). Microarray experiments were performed according to the protocols supplied by

Affymetrix as described previously in Schultz et al142

.

Statistical analysis

Statistical analyses were performed in GraphPad Prism 5 using Student t tests. A P value less than

0.05 were considered statistically significant. Unless otherwise indicated, means and standard

deviation are shown.

78

3.3.3 Results

Comparative analysis of OSCAR expression in myeloid cells

OSCAR is known to be expressed on myeloid cells134;136

, but comparative expression level has not

been reported. We analyzed OSCAR expression on the surface of DCs, monocytes and

macrophages derived from the same donor. Freshly isolated monocytes were analyzed while DCs

and macrophages were generated by culturing the monocytes for 6 days in GM-CSF and IL-4 or M-

CSF, respectively. Flow cytometry analysis showed that the expression levels of OSCAR differ

between these myeloid cells. OSCAR was expressed at the highest level on DCs as compared to

monocytes and macrophages. No difference in expression of OSCAR was observed between

monocytes and macrophages (Fig. 1)

Den

dritic

cel

ls

Monocy

tes

Mac

rophag

es

0

5

10

15

p=0.0001

p=0.1

p=0.005

OS

CA

R-P

E (

MF

I)

Fig. 1. DCs express higher level of OSCAR than monocytes and macrophages. Human monocytes were purified

and used either directly for analysis or for generating DCs and macrophages. Monocytes were differentiated into DCs

(50ng/ml GM-CSF and IL-4) or to macrophages (40ng/ml M-CSF) for 6 days. Cells were incubated with PE-conjugated

anti-human OSCAR mAbs or the corresponding isotype control for analysis of cell surface OSCAR expression. Data

are expressed as ∆ mean fluorescence intensity (∆MFI), the MFI minus the fluorescence detected with the isotype

control. Data for four independent donors are shown.

OSCAR-dependent ECM degradation by macrophages

The ability of macrophages to degrade ECM was tested using a Gelatin Invadopodia Assay. In this

set-up, gelatin represents not only a substrate for proteases, but also serves as an OSCAR ligand.

Monocytes from healthy controls were differentiated into macrophages. The macrophages were

transferred onto a glass coverslip coated with FITC-labelled gelatin in the presence or absence of

79

plg, with or without anti-OSCAR mAb or isotype control. After 24 hours their ability to degrade the

gelatin was measured by fluorescence microscopy as areas devoid of FITC-fluorescence.

We found that inhibiting OSCAR-collagen signaling led to reduced matrix degradation by

macrophages, which was not observed in the case of the control mAb (Fig. 2A). Exogenous Plg

enhanced gelatin degradation, which again was decreased when OSCAR was blocked (Fig. 2B).

This indicates a role of OSCAR in ECM degradation by macrophages. Aprotinin, which at low

concentrations works as a selective plasmin inhibitor, partially suppressed the macrophage-

mediated gelatin degradation. The same tendency was observed for all five donors analyzed when

quantifying the areas of degradation using the ImageJ software (Fig. 2C).

0

20

40

60

80

PLG

p=0.057

p=0.0039

ns

ns

Untreated

Isotype

Aprotinin

a-OSCAR

C

% d

eg

rad

ed

are

a

Fig. 2. Blocking of OSCAR signaling leads to reduced gelatin degradation by macrophages. Macrophages

generated from monocytes in the presence of 2000U M-CSF for six days were placed on FITC-labelled gelatin-coated

coverslips for 24 hours. All receiving 1000U/ml M-CSF during gelatin culture. They were cultured (A) without or (B)

with plasminogen (250nM) in the presence of either OSCAR-blocking mAb, OSCAR non-blocking mAb (1ug/ml) or

50nM Aprotinin. Top row pictures show merged F-actin (TRICT-Phalloidin, Red) and nuclei (DAPI, blue) stain; the

bottom row pictures show the FITC-labelled gelatin where degradation is visualized as areas devoid of FITC-staining

by fluorescent microscopy at 20x magnification. (C) The bars represent mean (n=5 donors, in 5 independent

experiments) areas±SD of degradation as % of total analyzed gelatin-coated area relative to untreated cell numbers.

OSCAR-collagen signaling does not mediate secretion of inflammatory mediators by

macrophages

80

Based on the data from the gelatin degradation assay, we hypothesized that OSCAR plays a role in

macrophage-mediated matrix degradation. Together with our findings in DCs and our in-house data

(preliminary, by Svetlana Panina, Novo Nordisk A/S), showing that cross-linking of OSCAR with

plate-bound mAbs results is a release of various pro-inflammatory cytokines, chemokines and

MMPs by macrophages, we attempted to look further into the mechanisms behind the OSCAR-

regulated matrix degradation. Macrophages were seeded in uncoated wells or wells coated with

ColII, in the presence or absence of the anti-OSCAR blocking mAb or isotype control. The

supernatants were analyzed for the presence of cytokines, chemokines and MMPs, as well as for the

MMP activity.

We analyzed a panel of cytokines, chemokines and MMPs in the supernatants of

macrophages cultured on ColII overnight, using Bioplex. We did not observe any effect of ColII on

the macrophage activation. In line, no change in secretion was seen when adding anti-OSCAR

mAbs or the isotype control (see representative cytokines and chemokines, and MMPs in Fig. 3A &

B.

Untr

eate

dCol

Col +

0.1

µg/m

l a-O

SCAR

Col +

1µg

/ml a

-OSC

AR

Col +

10µ

g/ml a

-OSC

AR

Col +

0.1

µg/m

l Iso

type

CoI +

1µg

/ml I

soty

pe

Col +

10µ

g/ml I

soty

pe

0

20

40

60

80

TN

F-

(p

g/m

l)

Untr

eate

dCol

Col +

0.1

µg/m

l a-O

SCAR

Col +

1µg

/ml a

-OSC

AR

Col +

10µ

g/ml a

-OSC

AR

Col +

0.1

µg/m

l Iso

type

CoI +

1µg

/ml I

soty

pe

Col +

10µ

g/ml I

soty

pe

0

5

10

15

20

RA

NT

ES

(p

g/m

l)

Untr

eate

dCol

Col +

0.1

µg/m

l a-O

SCAR

Col +

1µg

/ml a

-OSC

AR

Col +

10µ

g/ml a

-OSC

AR

Col +

0.1

µg/m

l Iso

type

CoI +

1µg

/ml I

soty

pe

Col +

10µ

g/ml I

soty

pe

0

5

10

15

GM

-CS

F (

pg

/ml)

Untr

eate

dCol

Col +

0.1

µg/m

l a-O

SCAR

Col +

1µg

/ml a

-OSC

AR

Col +

10µ

g/ml a

-OSC

AR

Col +

0.1

µg/m

l Iso

type

CoI +

1µg

/ml I

soty

pe

Col +

10µ

g/ml I

soty

pe

0

50

100

150

200IL

-12(p

40)

(pg

/ml)

A

81

Untr

eate

dCol

Col +

0.1

ug/ml a

-OSCAR

Col +

1ug/m

l a-O

SCAR

Col +

10u

g/ml a

-OSCAR

Col I

I + 0

.1ug/m

l a-T

NP

Col I

I + 1

ug/ml a

-TNP

Col I

I + 1

0ug/m

l a-T

NP

0

1000

2000

3000

4000

5000

MM

P2 (

pg

/ml)

Untr

eate

dCol

Col +

0.1

ug/ml a

-OSCAR

Col +

1ug/m

l a-O

SCAR

Col +

10u

g/ml a

-OSCAR

Col I

I + 0

.1ug/m

l a-T

NP

Col I

I + 1

ug/ml a

-TNP

Col I

I + 1

0ug/m

l a-T

NP

0

20

40

60

80

100

MM

P3 (

pg

/ml)

Untr

eate

dCol

Col +

0.1

ug/ml a

-OSCAR

Col +

1ug/m

l a-O

SCAR

Col +

10u

g/ml a

-OSCAR

Col I

I + 0

.1ug/m

l a-T

NP

Col I

I + 1

ug/ml a

-TNP

Col I

I + 1

0ug/m

l a-T

NP

0

20

40

60

80

100

MM

P8 (

pg

/ml)

Untr

eate

dCol

Col +

0.1

ug/ml a

-OSCAR

Col +

1ug/m

l a-O

SCAR

Col +

10u

g/ml a

-OSCAR

Col I

I + 0

.1ug/m

l a-T

NP

Col I

I + 1

ug/ml a

-TNP

Col I

I + 1

0ug/m

l a-T

NP

0

50000

100000

150000

MM

P9 (

pg

/ml)

B

Fig. 3. OSCAR-collagen interaction in macrophages does not lead to cytokine, chemokine or MMP release.

Human CD14+ cells were differentiated into macrophages by treatment with M-CSF for six days. The macrophages

were seeded on uncoated or Col-coated (10µg/ml) IWAKI plates. The macrophages were pre-incubated with either

0.1µg/ml, 1 µg/ml or 10µg/ml anti-OSCAR mAb or isotype control. After 20 hours, supernatants were analyzed, using

Bioplex. A) Analyzed cytokines/chemokines were IL-1, IL-6, IL-8, IL-10, IL-12(p40), IL-12(p70), GM-CSF, MCP-1,

MIP1, RANTES, TNF- and M-CSF. Only TNF-, RANTES, GM-CSF and IL-12(p40) are shown as representatives.

B) Analyzed MMPs were MMP1, MMP2, MMP3, MMP8, MMP9, MMP10 and MMP13. Only MMP2, MMP3, MMP8

and MMP9 are shown as representative. Mean concentrations ± SD of triplicates for one donor out of four independent

donors is shown.

MMPs are secreted as inactive pro-enzymes and undergo extracellular activation into the active

enzymes. Thus, to assess the relative potential of MMPs to degrade matrix, a more suitable method

would be to measure the enzyme activity. Since MMP secretion by macrophages was not affected

by collagen, we tested whether OSCAR-collagen signaling can influence MMP activity. We used a

fluorogenic substrate for several MMPs to determine the MMP activity. No difference in MMP

activity was observed between the untreated and collagen-treated macrophages. Correspondingly,

blocking OSCAR-collagen interaction did not influence MMP activity at the conditions used (Fig.

4).

82

Untr

eate

dCol

Col +

0.1

µg/m

l a-O

SCAR

Col +

g/ml a

-OSCAR

Col +

10

µg/m

l a-O

SCAR

Col +

0.1

µg/m

l iso

type

Col +

1 µ

g/ml i

soty

pe

Col +

10

µg/m

l iso

type

0

2.0106

4.0106

6.0106

8.0106

1.0107

RF

U a

t 360n

m

Fig. 4. Macrophages seeded on collagen do not activate MMPs. Human CD14+ cells were differentiated into

macrophages by treatment with M-CSF for six days. The macrophages were untreated or seeded on Col-coated

(10µg/ml) IWAKI plates. The macrophages were pre-incubated with 0.1µg/ml, 1 µg/ml or 10µg/ml anti-IIOSCAR mAb

or isotype control. After 20 hours, MMP activity in the supernatants was analyzed, using a Microplate reader Paradigm

at 360nm. Mean concentrations ± SD of triplicates wells for one donor out of four independent donors is shown.

OSCAR-Collagen signaling does not regulate gene expression in macrophages

Macrophages from six healthy donors were differentiated as described previously and stimulated

with collagen for four hours with or without anti-OSCAR mAbs or isotype control. The genome-

wide gene expression profiling was done using microarray in untreated cells and macrophages

treated with ColII in the presence or absence of anti-OSCAR mAb or isotype control mAb. The

analysis revealed that no genes were regulated by ColII treatment in macrophages in an OSCAR-

dependent manner. Examples of lack of gene regulation are shown in (Fig. 5) encompassing the top

six genes regulated in monocytes by OSCAR-collagen, however not affected in macrophages (Fig.

5).

83

Fig. 5. Stimulation with collagen does not regulate gene expression in macrophages. Human CD14+ cells were

differentiated into macrophages by treatment with M-CSF for six days. The macrophages were seeded on uncoated or

Col-coated (10µg/ml) IWAKI plates. The macrophages were pre-incubated with 1 µg/ml anti-OSCAR mAb or isotype

control. After 4 hours, total cellular RNA was isolated, and genome-wide RNA levels were determined using

Affymetrix GeneChip microarray analysis. A false discovery rate at 5% and a biological criterion of minimum two-fold

change was used in the analysis. Data are presented on a log2-scale. Samples from six independent donors were

analyzed.

As illustrated by the Principal Component Analysis plot below, the most pronounced effect was

donor-to-donor variation. When grouped by treatment no distinct correlation was seen (Fig. 6).

Fig. 6. Principal Component Analysis plot. A) shows data grouped by donors. B) shows data grouped by treatment.

84

3.3.4 Discussion

Macrophages are known to be crucial for synovial inflammation. They exert a variety of immuno-

regulatory, phagocytic and matrix degrading functions. In the joint of RA patients, macrophages are

found to secrete multiple proinflammatory cytokines, as well as proteases that are involved in the

degradation of ECM5;6

. In a variety of proof-of-principle clinical studies, it has been shown that the

number of macrophages infiltrating synovial tissue correlates with disease severity and inversely –

with therapeutic benefit, independently of the type of therapy153

. It has been proposed that reduction

of the macrophage numbers after successful treatment does not result from cell death, but rather

from the egress of macrophages from synovial tissue and/or prevention of the influx from

circulation196;197

. Therefore, it has been suggested that targeting monocyte/macrophage

migration/invasion, survival and activation/differentiation can offer a therapeutic potential.

Here, we have observed that gelatin degradation by macrophages was OSCAR

dependent. In our initial experiments, cross-linking of OSCAR in macrophages led to the secretion

of cytokines, chemokines and MMPs (unpublished, S. Panina, Novo Nordisk A/S), indicating that

OSCAR signaling is functional in this cell type. Hence, OSCAR engagement by its native ligand

collagen was thought to be able to activate macrophages. However, we have not been able to detect

any effect on the regulation of gene expression or on the cytokine, chemokine and MMP secretion,

as well as MMP activity when culturing macrophages on collagen. Since activation of other cell

types (i.e. DCs and monocytes) derived from the same donor was drastic, the absence of response in

macrophages to collagen is difficult to explain based on our observations, and would require

additional experiments. The observed discrepancies between responses of the macrophages and

DCs/monocytes to collagen may, at least in part, be attributed to the different expression levels of

various immunoreceptors, including OSCAR expression, or the components of their signaling

machinery. We have analyzed the expression level of OSCAR in monocytes, macrophages and DCs

derived from the same donors. The cell surface expression of OSCAR in monocytes and

macrophages was comparable and significantly lower than that in DCs. We have previously shown

that OSCAR-collagen interaction activates monocytes142

, so the lower expression level in

macrophages compared to DCs may not account for the entire absence of collagen responsiveness

in macrophages.

Collagens are known to interact with multiple immune cell receptors, including

adhesion molecules (integrins) and non-integrin receptors99

. Integrins may synergize with ITAM

receptors providing activating signals to the cells99

. The non-integrin group includes inhibitory

85

ITIM-type receptors, which are known to suppress intracellular signaling by ITAM receptors, hence

regulating the immune response by balancing activating and inhibitory signals198

. The ITIM

receptor LAIR-1 was recently shown to interact with ColI, -II and -III128;199;200

. The LAIR-1 binding

sites on collagen were found not to overlap with OSCAR binding sites on collagen99

. In-house data

on LAIR-1 expression showed that LAIR-1 was expressed at moderate level on monocytes, at low

level in DCs and was highly expressed in differentiated macrophages (unpublished data by S.

Panina and L. Bentzon, Novo Nordisk A/S). LAIR-1 may potentially inhibit OSCAR activation

when both receptors are engaged by collagen. It is possible that the ratio of inhibitory and activating

signals may define the functional outcome of cell stimulation by collagen. LAIR-1 may therefore

contribute to the lack of the effect of collagen seen in macrophages. One may further speculate that

additional ITIM-receptors could interfere with OSCAR-signaling. The inhibitory receptor leucocyte

Ig-like receptor-1/Ig-like Transcript 2 (CD85j) has been shown to negatively regulate OSCAR

induced survival and cytokine production in DCs12

. A better mechanistic understanding of OSCAR-

collagen signaling in macrophages could be explored by inhibiting ITIM-signaling. This could be

performed using pharmacological inhibitors of ITIM signaling components or by knock-down of

individual ITIM receptors, or by using specific receptor-blocking antibodies. Moreover, the

interfering signals from other collagen receptors could also be avoided by using specific OSCAR-

binding collagen peptides instead of full-length collagen. If collagen still cannot induce activation in

macrophages when ITIM-signaling is blocked, another possibility is that OSCAR requires a co-

receptor for full activation. Such putative co-receptor might be expressed on monocytes and DCs,

but not on macrophages. Merck and colleagues have shown that the effect of OSCAR ligation in

neutrophils is weaker compared to the effect observed in monocytes134

. They speculated that it

could be explained by a limited availability of FcRγ in neutrophils, which serves as signaling

subunit for multiple receptors expressed and functioning in immune cells.

Despite the lack of effect of collagen on the activation of macrophages, we have shown that anti-

OSCAR mAb inhibits degradation of gelatin by macrophages. The effect was reproducible and seen

in at least five independent experiments. Gelatin has a high content of collagen that, besides serving

as a substrate for macrophage degradation, functions as a ligand for OSCAR. Gelatin degradation is

primarily mediated by MMPs and various proteases, like plasmin201

. We found that macrophage

degradation of gelatin was greatly potentiated by the addition of plasminogen. When OSCAR was

blocked at these conditions, the degradation was inhibited to the level observed when no

86

plasminogen was added. This suggests that OSCAR-signaling induce factors important for plasmin

activation. We have previously identified several genes involved in ECM turnover being regulated

by OSCAR-signaling in DCs and monocytes. These genes include several MMPs and urokinase-

type plasminogen activator receptor (uPAR)142

. We speculate that OSCAR signaling may

potentially control ECM degradation via regulation of MMP and conversion of plasminogen to

plasmin. Unfortunately, I did not study the mechanisms of OSCAR involvement in macrophage

mediated degradation of gelatin further during my stay at the University of Melbourne.

We have tried to clarify the mechanisms of macrophage-mediated gelatin breakdown

by using a more simple system, namely a purified collagen, as the natural ligand of OSCAR. Even

though gelatin contains collagen in its polymerized form, the composition of the two substrates used

in our study is very different and this fact may be the major reason why collagen did not promote

activation of macrophages while gelatin did. Besides collagen, other ECM components are present

in gelatin, which may interact with ECM receptors and potentiate the effect of the OSCAR

signaling. Alternatively, binding sites for inhibitory (e.g. ITIM) receptors might not be accessible in

gelatin. Studies have shown that various matrix proteins are crucial for macrophage MMP

secretion202-204

. It has been shown that macrophages cultured on ColI and ColII but not ColIV,

laminin, fibronectin and elastin can induce collagenase (MMP-1, -8 and -13) production202

. Lepidi

et al. reported that MMP synthesis by macrophages was decreased on polymerized ColI and slightly

increased on monomeric collagen203

. Moreover, in a study searching for immune-modulating

biomaterial to regulate macrophage activity, Franz et al. found that ECM composed of ColI and

high-sulfated hyaluronan dampened the activity of pro-inflammatory macrophages204

. Taken

together, these reported data indicate that the nature of ECM may drastically influences macrophage

activity. Further experiments should be performed to mechanistically assess activation of the

OSCAR signaling in macrophages using gelatin as a surrogate ligand.

In summary:

1. OSCAR is expressed on macrophages and can potentially be functional, as shown by cross-

linking studies

2. Purified ColII does not stimulate macrophage activation at the conditions used

3. OSCAR signaling may control macrophage-mediated degradation of collagen-rich ECM,

gelatin, possibly by regulating MMP secretion and/or activation and converting

plasminogen into plasmin

87

4. Discussion

It is well established that OSCAR plays a role in osteoclastogenesis, as it has been shown in in vivo

and in vitro studies29;32;137

. In humans, OSCAR is more widely expressed than in mice, where the

expression is restricted to osteoclasts29

. Before collagen was discovered as the ligand of

OSCAR137;142

, the functions of OSCAR in immune cells were studied by cross-linking OSCAR

with mAbs11;134;136

. Therefore, the main goal of this PhD project was to investigate the function of

OSCAR following interaction with its natural ligand, collagen, in human immune cells.

OSCAR is constitutively expressed on human myeloid cells. Consequently, the

availability of the OSCAR ligand will most likely determine the activation of OSCAR. OSCAR

plays an important role in OC function and in the maintenance of bone205

. However, in immune

cells the function of OSCAR is still not yet elucidated. The group of Trowsdale speculated that

collagen could be regarded as a “damage-associated molecular pattern” that could be sensed by

OSCAR on immune cells to detect alterations in the ECM137

. This would be beneficial at the

homeostatic conditions, but could potentially be harmful in chronic inflammatory situations.

Collagens are recognized to play a major role in RA. From animal models we know

that collagen can induce arthritis (collagen-induced arthritis, CIA, models). Moreover, collagen

fragments and antibodies against collagens have been identified in the sera and synovial fluid of

some RA patients 158;159;206;207

. It has further been suggested that the severity of RA may correlate

with an increase in collagen degradation190

. In RA, it is likely that collagen, which is normally

embedded in the ECM or under the layer of endothelial cells, chondrocytes and fibroblasts,

becomes exposed due to cellular reorganization and ECM remodeling within the RA joint43;104;137

.

OSCAR has been demonstrated to be expressed on OCs and mononuclear cells in the joint tissue of

RA patients and being up-regulated on the circulating monocytes141

. We have speculated that the

collagen exposure upon ECM remodeling in the RA joints and the concurrent influx of

inflammatory OSCAR-expressing immune cells like monocytes, macrophages and DCs, may

significantly contribute to the RA disease pathology. This was supported by the findings by

Trowsdale and colleagues who have identified OSCAR-expressing mononuclear cells in contact

with collagen type I and II at the bone remodeling sites137

. It is therefore crucial to gain further

knowledge about the functional outcome of the OSCAR-collagen interaction in immune cells, as a

way to discover new therapeutic interventions.

88

During the course of inflammation, monocyte-derived DCs (mo-DCs) may develop in the inflamed

tissue from blood monocytes208;209

. Due to the relevance in RA, we chose this DCs type to study the

effect of OSCAR-collagen interaction. In paper I we describe the identification of ColI and ColII

as the ligands of OSCAR and investigate the functional outcome of the OSCAR-collagen

interaction in DCs142

. Monocytes and macrophages are central to the inflammation observed in

RA5. They massively infiltrate the joint, have an activated phenotype and produce high levels of

TNF-6;151;152;210

. When entering the joint, monocytes become further activated upon adherence to

ECM and exposure to the complex inflammatory environment encountered211

. The functional

outcome of monocyte activation by ECM would provide useful information about the pathogenesis

of RA. Paper II primarily describes the activation of monocytes by the OSCAR-collagen

interaction142

. Macrophages are a major source of cytokines and chemokines, and degrading

enzymes which drive inflammation and joint destruction6. Treatments which have been shown to

improve the disease activity led to a decrease in macrophage numbers in the joint153;194

. Based on

the findings in papers I and II, part III of the result section assess the role of OSCAR in

macrophages using two different substrates: ColII and collagen-rich gelatin. Have

A thorough discussion of the results is provided in the Discussion section of the included paper I,

paper II and part III. Hence, the section below covers a combined discussion with a focus on the

mechanisms and biology of OSCAR-collagen interaction in DCs, monocytes and macrophages,

related to the RA pathology.

The identification of collagen as a ligand for OSCAR was done prior to the start of my PhD project

by the group of Dr. Panina (Novo Nordisk A/S) in parallel with the group of Trowsdale, with the

latter being the first to publish the data137

. Both ColI and ColII were found to bind OSCAR (Paper

I) and initially both collagens were used to stimulate the DCs. Our results showed that ColI and

ColII elicited the same effect in DCs. However, ColII was more potent than ColI, as ColII generated

a greater cytokine response and induced higher expression of surface activation markers than ColI.

When comparing the gene profiles of ColI- and ColII-stimulated DCs, we saw that over 90% of the

genes overlapped. We therefore chose to continue the studies using ColII as our primary OSCAR

ligand.

89

To understand the extent of the involvement of OSCAR in DC biology, we investigated the effect

of adding anti-OSCAR-blocking mAb while generating iDCs from blood monocytes. We found that

adding anti-OSCAR mAb did not have an effect on the differentiation of monocytes to iDCs under

the conditions used in these studies. This suggests that iDC differentiation does not require

OSCAR-collagen signaling. Further, to exclude any indiscriminate inhibitory effect on DC

activation by anti-OSCAR mAb, we evaluated the effect of anti-OSCAR mAb on LPS activated

DCs. We found that anti-OSCAR mAb did not have an effect on the LPS-driven cytokine release.

Hence, we believe that adding OSCAR mAb has no effect on DCs when collagen is not present.

As demonstrated in paper I and II, we had several main findings when exploring the outcome of

the OSCAR engagement by collagen in DCs and monocytes. Culturing DCs and monocytes on

collagen led to an enhanced secretion of a variety of inflammatory markers, including chemokines

and cytokines, in an OSCAR-dependent manner. Chemokine and cytokines are important in the

orientation and amplification of the adaptive immune response212

. Among the chemokines detected,

IL-8 and RANTES were found at high levels. These chemokines recruit immune cells, including T

cells, neutrophils, granulocytes and monocytes to the joints. These cells further mediate the

generation of reactive oxygen species, promote angiogenesis, and modulate the function of other

immune cells213;214

. Merck and colleagues suggested that the role of OSCAR in DCs was not

proinflammatory since they could not detect any or low amounts of cytokines when cross-linking

OSCAR on DCs or monocytes11;134

. In contrast, we observed a drastic induction of proinflammatory

cytokines, like TNF-, IL-1 and IL-6, by DCs and monocytes when cultured on collagen,

suggesting that OSCAR on DCs and monocytes have a proinflammatory function. Additionally, we

found that collagen, in an OSCAR dependent manner, induced secretion of TNF- and IL-8 by

synovial fluid cells from RA patients, showing that the OSCAR pathway is active in immune cells

present in the RA synovium. TNF-, IL-1, IL-6 and IL-8 may all be found at high levels in the RA

joints and seem to play a crucial role in the course of the disease25

. The pro-inflammatory cytokines

contribute to the RA pathology in several ways. They stimulate osteoclastogenesis, increase the

bone-resorbing capacity of osteoclasts, stimulate secretion of MMPs and nitric oxide and prevent

new synthesis of collagens and proteoglycans, hence contributing to bone erosion and cartilage

degradation43;104;113

.

90

To evaluate the underlying molecular mechanisms of the OSCAR-collagen mediated

cell activation, we analyzed the global gene expression profile upon OSCAR engagement in DCs

and monocytes. OSCAR engagement by collagen up-regulated expression of genes involved in the

mitogen-activated protein kinases (MAPK) pathways, like p38 and ERK, as well as PI3K and the

NFB family of transcription factors. Presumably as a result of the activation of NFB and MAPK

signaling, several pro-inflammatory genes, which are known to be transcriptional targets of these

pathways, were strongly up-regulated. Of importance in DCs, OSCAR-collagen interaction

regulated the expression of genes involved in the immune regulation and activation. Among these,

the most up-regulated genes were TNF-, IL-1 and IL-1, together with several chemokines

including CCL20, CCL5/RANTES, CXCL8/IL-8 as well as chemoattractants for endothelial

precursor cells CXCL1, -2 and -3. In monocytes, expression of genes which products are involved

in the immune regulation and cell activation were also induced by OSCAR-collagen interaction. Of

interest was the up-regulation of genes coding for TNF-, IL-1, CCL24, CCL20, CCL7, CXCL2

and -3, which are essential for amplification of the immune response and attracting lymphocytes,

monocytes, macrophages and endothelial cells25;44

. For both DCs and monocytes, genes involved in

cell-matrix interaction, such as osteopontin, syndecans and MMPs were found to be up-regulated.

The gene profile analysis supports our findings on the protein level that OSCAR signaling can

activate monocytes and DCs to secrete proinflammatory mediators.

In the absence of growth factors, most cells would undergo apoptosis. Paper I and II make it

evident that OSCAR-collagen interaction protects monocytes and DCs from apoptosis induced by

the absence of growth factors. Our results are in line with the findings made by other groups,

showing the pro-survival effects of OSCAR11;134;147

. Based on our microarray analysis one may

speculate about the mechanism by which OSCAR-collagen interaction promotes survival of DCs

and monocytes. The gene profile analysis of DCs and monocytes revealed that several pro- and anti-

apoptotic pathways were affected, like the MAPK and the TNF related weak inducer of apoptosis

(TWEAK) signaling. However, the specific mechanisms behind the increased viability appear

different for the DCs and monocytes. OSCAR-collagen interaction in DCs seems to protect them

from apoptosis by up-regulating anti-apoptotic genes, such as BCL2A1 and multiple components of

the MAPK/ERK pathway, and by down-regulating members of the caspase family, such as caspase

6 and 8. Fewer genes controlling cell survival were found to be regulated in monocytes by OSCAR.

Nevertheless, the death receptor signaling pathway promoting apoptosis was seen to be down-

91

regulated (TNFSF10/TRAIL), while TNFRSF4/CD134 which is involved in suppressing apoptosis

was up-regulated by Col in an OSCAR-dependent manner.

Additionally, we have shown that both DCs and monocytes produce TNF- when

stimulated with collagen, which in an autocrine manner could contribute to the increased

viability169;215;216

. The increased viability of DCs and monocytes could ensure a long-term activity,

sustaining the inflammation seen in the joints of RA patients.

DCs are potent APC cells that are crucial for the communication between the innate and adaptive

immune system. After Ag capture, iDCs migrate to the lymph nodes and the spleen, where they

undergo maturation and present their antigen-peptide complexes to T cells. In RA, DCs are found in

the synovial tissue and synovial fluid, where they may play a role of powerful immune stimulators.

In the RA synovial tissue, DCs have been found within lymphoid aggregates expressing a variety of

co-stimulatory receptors and being capable of activating T cells57;217

. Paper I shows that upon

OSCAR-collagen interaction the DCs increase the expression of numerous phenotypic markers for

cellular activation and maturation, such as CD83, CD86 and HLA-DR. Moreover, several

components of the immune synapse including CD40, CD80, CD83, CD86 and ICAM-1, were found

to be transcriptional targets of the OSCAR-collagen signaling. The interaction of DCs with T cells

is central for the inflammation seen in RA57

. When DCs mature they become capable of inducing

clonal expansion of T cells and the cytokines secreted by the DCs can skew the nature of the T cell

responses217

. The Col-DCs showed a significant difference in their ability to induce T cell

proliferation in the DC-T cell co-culture studies compared to the control iDCs. When adding an

anti-OSCAR blocking mAb during DC maturation, these DCs had a reduced T cell stimulatory

capacity. The T cells that were co-cultured with the Col-DCs produced high amounts of

proinflammatory cytokines and chemokines. We found both Th1 and Th2 cytokines in the

supernatants of the Col-DCs-T cell co-cultures. We did not see any clear polarization of the T cells,

as the cytokine levels did not seems to be skewed. The generated T cells may further contribute to

bone remodeling by expressing RANKL hence increasing osteoclastogenesis more.

Part III of my PhD work describes the role of OSCAR-collagen interaction in macrophages. We

and others136

have found OSCAR to be expressed on human DCs, monocytes and macrophages.

When analyzing the OSCAR expression levels on these cells, we identified a higher expression of

OSCAR on DCs compared to monocytes and macrophages. In-house data by the group of Dr.

92

Panina have shown that OSCAR may potentially be functional in macrophages as cross-linking of

OSCAR enhanced the secretion of cytokines and MMPs (preliminary data by Dr. Panina, Novo

Nordisk A/S). However, in macrophages the activity of OSCAR seems to be different from DCs

and monocytes, as we did not observe any collagen-mediated effects in the analyzed macrophages.

We speculated that the unresponsiveness could partly be due to the lower OSCAR expression than

that in DCs in combination with a high expression of the inhibitory collagen receptor LAIR1, or

maybe a defective capacity of the OSCAR-signaling machinery. However, when assessing

degradation of gelatin by macrophages, we found that blocking OSCAR effectively decreased the

degradation. Hence, indicating that OSCAR plays a role in enhancing macrophage-mediated gelatin

degradation. Gelatin degradation is primarily mediated by MMPs and other proteases such as

plasmin. Through gene analysis in DCs and monocytes we have found that OSCAR-signaling

affects genes involved in the ECM remodeling, including MMPs and the uPA receptor142

. When

plasminogen was present, matrix degradation was more efficient; however blocking OSCAR-

signaling inhibited the degradation to the level observed when no plasminogen was added. This

indicates that OSCAR signaling induces factors important for plasmin activation. Hence, the

mechanisms of OSCAR-regulated gelatin degradation might be MMP and plasmin dependent.

Altogether, OSCAR engagement in macrophages might contribute to the invasion of macrophages

into the synovium and to joint destruction in RA.

We have shown that gelatin but not collagen had an OSCAR-dependent effect serving

as a ligand for OSCAR. Although the substrates are related there is a clear difference in structural

composition and preparation of ColII and gelatin. Gelatin is a mixture of several collagens, mainly

ColI and III, and other ECM components like elastin, fibrillin and fibronectin are present98;106;218-220

.

The ColII used in my experiments is relatively pure (90%), with only little contribution of other

collagens. The OSCAR and LAIR1 binding sites may therefore be different between ColII and

gelatin. The gelatin experiments were conducted during my stay in Australia and unfortunately due

to a time limitation I did not study the mechanisms behind my observations. It remains to be

investigated what is causing the difference in the OSCAR-collagen effects between macrophages

and DCs/monocytes. One strategy could be to evaluate the contribution of ITIM-receptors, which

could be investigated by using specific receptor-blocking antibodies, inhibitors of ITIM signaling

components or knocking-down individual ITIM receptors. The interfering signals from other

collagen receptors could also be avoided by using specific OSCAR-binding collagen peptides

instead of full-length collagen. Finally, there is a possibility that OSCAR requires a putative co-

93

receptor for full activation, which might be expressed on monocytes and DCs, but not on

macrophages.

Figure. 10. Role of OSCAR in rheumatoid arthritis. Potential outcome of OSCAR-collagen interaction in the RA

joint is indicated in red boxes. OSCAR is a co-stimulator of osteoclastogenesis. In monocytes, DCs and macrophages

OSCAR shows immune-modulatory effects, including increased viability, promotion of activation, maturation (leading

to T cell expansion and polarization) and ECM degradation. The secreted cytokines and chemokines recruit more

immune cells, enhance cell stimulation and stimulate the release of proteases. Altogether, inflammation, and increased

osteoclastogenesis, mediates bone and cartilage degradation. Blocking of OSCAR signaling in the synovial tissue may

inhibit macrophage invasion, dampen DC-driven local T cell expansion and hence cytokine level and inhibit cartilage

and bone damage.

94

5. Conclusion

This PhD thesis demonstrates that collagen is a functional ligand of OSCAR in DCs and

monocytes. Our data emphasizes the importance of ECM remodeling in chronic inflammatory

conditions, as the exposed ECM interacts with immune cells contributing to the inflammation seen

in the joints of RA patients. We have shown that OSCAR-collagen interaction in DCs leads to an

increased viability, enhanced secretion of pro-inflammatory cytokines and chemokines, augmenting

the recruitment of immune cells to the joint. OSCAR promotes maturation of DCs, which leads to

expansion and activation/polarization of T cell characterized by cytokine secretion, thus further

enhancing the inflammatory response. We found that OSCAR promotes cell survival and activation

of monocytes. Monocytes were shown to secrete inflammatory cytokines and chemokines and to

up-regulate surface markers when stimulated with collagen in an OSCAR-dependent manner.

Furthermore, OSCAR was expressed and shown to be functional on synovial fluid cells from RA

patients. Surprisingly, the OSCAR-collagen pathway was not capable of activating macrophages.

Though, preliminary data indicate that OSCAR can be involved in the control of degradation of

collagen-rich gelatin by macrophages, the mechanisms controlling this effect need further

investigation. Altogether, our studies indicate that OSCAR is not only involved in

osteoclastogenesis, but may activate the immune response, hence contributing to the pathogenesis

of RA on multiple levels. This underlines the possibility that inhibition of OSCAR signaling may

halt both, the inflammatory response and the bone erosion associated with RA.

95

6. Perspectives

Current treatment options for RA are to reduce inflammation and delay joint destruction. The

simplified RA treatment paradigm relies on steroids and Non-steroidal anti-inflammatory drug

(NSAIDS) for mild RA, DMARDS for moderate RA and biologics, like TNF- inhibitors, for

severe RA, which all primarily target inflammation. Unfortunately, a large number of the patients

fail to respond to the therapeutics and the therapies can cause undesired side effects221;222

. Recently,

myeloid cells have come into focus as potential targets for RA treatment. Anti-RANKL mAbs have

been developed and demonstrated a favorable bone-protecting effect, but no effect on inflammation

in RA222

. So, even though new therapies were developed in the recent years, there is still a strong

need for better treatment options. Novo Nordisk A/S therefore explored the option of targeting

OSCAR for treatment of RA. The expectations were that targeting OSCAR would inhibit OCs

formation as well as myeloid cell activation to limit both bone erosion and inflammation.

6.1 OSCAR project

The purpose of this PhD was to investigate the function of OSCAR on immune cells when engaged

by its natural ligand collagen. A thorough examination and understanding of OSCAR biology have

been achieved in DCs and monocytes. Certainly, more experiments could be conducted. It could be

interesting to assess the contribution of collagen signaling through OSCAR in the presence of TLR

ligands or other inflammatory mediators, to evaluate any synergistic effects.

A better understanding of OSCAR-collagen signaling in macrophages could be

achieved. The interplay between ITAM and ITIM receptors would need to be further examined.

This could be accomplished by working at conditions were ITAM signaling is absent, for example

by inhibiting ITIM-signaling or by using specific OSCAR-binding peptides. Additional mechanistic

studies using gelatin as a ligand for OSCAR could be conducted to assess macrophage activation.

If a putative OSCAR co-receptor exists, it could be identified by co-

immunoprecipitation using anti-OSCAR Ab followed by mass-spectrometry.

Several studies have shown high plasticity of myeloid cells. Mo-DCs are seen to be

capable of transdifferentiating into bone-resorbing OCs9. OSCAR might play a role in this

transdifferentiation of iDCs to OCs. Long term culture of DCs on collagen could be conducted and

examined for the presence of multinucleated OCs expressing cathepsin K and TRAP (cytostaining)

96

and if such cells are found in the culture – continue further with functional assays, where the cells

will be examined for the ability to degrade bone slices.

In addition, it would be interesting to assess the ability of the Col-DCs activated T

cells to promote osteoclastogenesis from OCs-precursor cells. This could be tested in an

osteoclastogenesis assay using the T cells in co-culture with CD14+ cells in the presence or absence

of RANKL/M-CSF and/or RA synovial fluid.

6.2 Atherosclerosis

Another relevant indication for the use of anti-OSCAR mAbs could be in the inflammatory disease

atherosclerosis. The atherosclerotic lesions (especially, non-occlusive vulnerable plaques) contain

large numbers of activated immune cells, including monocytes, macrophages and DCs223-225

. The

immune cells are found to be located mainly at the rupture prone areas, increasing the risk of

thrombus formation, heart attacks and strokes226

. High levels of MMPs and oxidants are found

within the lesions causing ECM remodeling and protein modifications, which would make

collagens accessible for interaction with OSCAR expressed on the immune cells. When analyzing

target genes and pathways regulated by OSCAR-collagen signaling in DCs and monocytes, one of

the main regulated pathways was predicted to be “atherosclerosis signaling”. Of interest, one of the

most up-regulated genes in both DCs and monocytes was the LPL gene, coding for Lipoprotein

Lipase. LPL mediates lipoprotein uptake and dysregulation has been linked to disorders of

lipoprotein metabolism. Altogether, suggesting that OSCAR could play a role in the pathogenesis of

atherosclerosis.

6.3 Limitations to the OSCAR project

OSCAR is an immune receptor broadly expressed on myeloid cells in human, while expression is

restricted to OCs in mice and several other animals except higher primates (In-house data from

Novo Nordisk A/S)29;136

. In vitro data indicate that targeting of OSCAR may be efficient for

dampening inflammation and bone erosion. However, the difference in the expression pattern

makes it difficult to conduct in vivo proof-of-principle studies using animal disease models. Novo

Nordisk A/S has addressed the function of OSCAR on bone erosion and in osteoclastogenesis in

mouse models of inflammatory bone destruction (Collagen-Induced Arthritis (CIA) model and

Delayed-Type Hypersensitivity (DTH) model). Unfortunately, blocking OSCAR did not show any

97

significant effect on bone erosion, and OC formation and activity in the animal models examined.

The lack of efficacy may be explained by the fact that OSCAR is not expressed on peripheral

monocytes which seem to serve as precursors of OCs in the CIA and DTH models. The murine

bone marrow derived OCs precursors, even though they express OSCAR, may not contribute

significantly to the inflammatory bone loss. Thus, neither CIA nor DTH murine models are optimal

for testing OSCAR contribution to bone destruction. The OSCAR project group came up with

several proposals for addressing OSCAR function in other in vivo setups. The only species

expressing OSCAR in a similar to human pattern are monkeys. However, for ethical reasons Novo

Nordisk A/S does not conduct intervention studies in monkeys. It has been proposed to generate a

humanized mouse model (knock-in), where OSCAR would have the same expression pattern as in

humans, and/or mechanistically address OSCAR mode of action in a small clinical study.

Nevertheless, the management board decided to discontinue the OSCAR project due to

unpredictability of the outcome and high costs.

98

7. References

Reference List

1. Arron JR, Choi Y. Bone versus immune system. Nature 2000;408:535-536.

2. Rauner M, Sipos W, Pietschmann P. Osteoimmunology. Int.Arch.Allergy Immunol.

2007;143:31-48.

3. Takayanagi H. New developments in osteoimmunology. Nat.Rev.Rheumatol. 2012;8:684-

689.

4. Auffray C, Sieweke MH, Geissmann F. Blood monocytes: development, heterogeneity, and

relationship with dendritic cells. Annu.Rev.Immunol. 2009;27:669-692.

5. Gordon S, Taylor PR. Monocyte and macrophage heterogeneity. Nat.Rev.Immunol.

2005;5:953-964.

6. Kinne RW, Stuhlmuller B, Burmester GR. Cells of the synovium in rheumatoid arthritis.

Macrophages. Arthritis Res.Ther. 2007;9:224.

7. Satpathy AT, Wu X, Albring JC, Murphy KM. Re(de)fining the dendritic cell lineage.

Nat.Immunol. 2012;13:1145-1154.

8. Alnaeeli M, Park J, Mahamed D, Penninger JM, Teng YT. Dendritic cells at the osteo-

immune interface: implications for inflammation-induced bone loss. J.Bone Miner.Res.

2007;22:775-780.

9. Rivollier A, Mazzorana M, Tebib J et al. Immature dendritic cell transdifferentiation into

osteoclasts: a novel pathway sustained by the rheumatoid arthritis microenvironment. Blood

2004;104:4029-4037.

10. Brand U, Bellinghausen I, Enk AH et al. Influence of extracellular matrix proteins on the

development of cultured human dendritic cells. Eur.J.Immunol. 1998;28:1673-1680.

11. Merck E, de Saint-Vis B, Scuiller M et al. Fc receptor gamma-chain activation via hOSCAR

induces survival and maturation of dendritic cells and modulates Toll-like receptor

responses. Blood 2005;105:3623-3632.

12. Tenca C, Merlo A, Merck E et al. CD85j (leukocyte Ig-like receptor-1/Ig-like transcript 2)

inhibits human osteoclast-associated receptor-mediated activation of human dendritic cells.

J.Immunol. 2005;174:6757-6763.

13. Sallusto F, Schaerli P, Loetscher P et al. Rapid and coordinated switch in chemokine

receptor expression during dendritic cell maturation. Eur.J.Immunol. 1998;28:2760-2769.

14. Cyster JG. Leukocyte migration: scent of the T zone. Curr.Biol. 2000;10:R30-R33.

99

15. Bouchon A, Hernandez-Munain C, Cella M, Colonna M. A DAP12-mediated pathway

regulates expression of CC chemokine receptor 7 and maturation of human dendritic cells.

J.Exp.Med. 2001;194:1111-1122.

16. Gallo PM, Gallucci S. The dendritic cell response to classic, emerging, and homeostatic

danger signals. Implications for autoimmunity. Front Immunol. 2013;4:138.

17. Grakoui A, Bromley SK, Sumen C et al. The immunological synapse: a molecular machine

controlling T cell activation. Science 1999;285:221-227.

18. Bromley SK, Burack WR, Johnson KG et al. The immunological synapse.

Annu.Rev.Immunol. 2001;19:375-396.

19. Brownlie RJ, Zamoyska R. T cell receptor signalling networks: branched, diversified and

bounded. Nat.Rev.Immunol. 2013;13:257-269.

20. Kaech SM, Cui W. Transcriptional control of effector and memory CD8+ T cell

differentiation. Nat.Rev.Immunol. 2012;12:749-761.

21. Noble BS, Stevens H, Loveridge N, Reeve J. Identification of apoptotic changes in

osteocytes in normal and pathological human bone. Bone 1997;20:273-282.

22. Gu G, Mulari M, Peng Z, Hentunen TA, Vaananen HK. Death of osteocytes turns off the

inhibition of osteoclasts and triggers local bone resorption. Biochem.Biophys.Res.Commun.

2005;335:1095-1101.

23. Gori F, Hofbauer LC, Dunstan CR et al. The expression of osteoprotegerin and RANK

ligand and the support of osteoclast formation by stromal-osteoblast lineage cells is

developmentally regulated. Endocrinology 2000;141:4768-4776.

24. Teitelbaum SL. Bone resorption by osteoclasts. Science 2000;289:1504-1508.

25. McInnes IB, Schett G. Cytokines in the pathogenesis of rheumatoid arthritis.

Nat.Rev.Immunol. 2007;7:429-442.

26. Boyle WJ, Simonet WS, Lacey DL. Osteoclast differentiation and activation. Nature

2003;423:337-342.

27. Teitelbaum SL, Ross FP. Genetic regulation of osteoclast development and function.

Nat.Rev.Genet. 2003;4:638-649.

28. Lacey DL, Timms E, Tan HL et al. Osteoprotegerin ligand is a cytokine that regulates

osteoclast differentiation and activation. Cell 1998;93:165-176.

29. Kim N, Takami M, Rho J, Josien R, Choi Y. A novel member of the leukocyte receptor

complex regulates osteoclast differentiation. J.Exp.Med. 2002;195:201-209.

30. Kim K, Kim JH, Lee J et al. Nuclear factor of activated T cells c1 induces osteoclast-

associated receptor gene expression during tumor necrosis factor-related activation-induced

cytokine-mediated osteoclastogenesis. J.Biol.Chem. 2005;280:35209-35216.

100

31. Mocsai A, Abram CL, Jakus Z et al. Integrin signaling in neutrophils and macrophages uses

adaptors containing immunoreceptor tyrosine-based activation motifs. Nat.Immunol.

2006;7:1326-1333.

32. Koga T, Inui M, Inoue K et al. Costimulatory signals mediated by the ITAM motif

cooperate with RANKL for bone homeostasis. Nature 2004;428:758-763.

33. Takahashi N, Akatsu T, Udagawa N et al. Osteoblastic cells are involved in osteoclast

formation. Endocrinology 1988;123:2600-2602.

34. Nakashima T, Takayanagi H. The dynamic interplay between osteoclasts and the immune

system. Arch.Biochem.Biophys. 2008;473:166-171.

35. Simonet WS, Lacey DL, Dunstan CR et al. Osteoprotegerin: a novel secreted protein

involved in the regulation of bone density. Cell 1997;89:309-319.

36. Zupan J, Jeras M, Marc J. Osteoimmunology and the influence of pro-inflammatory

cytokines on osteoclasts. Biochem.Med.(Zagreb.) 2013;23:43-63.

37. Cheng J, Liu J, Shi Z et al. Interleukin-4 inhibits RANKL-induced NFATc1 expression via

STAT6: a novel mechanism mediating its blockade of osteoclastogenesis. J.Cell Biochem.

2011;112:3385-3392.

38. Takayanagi H, Ogasawara K, Hida S et al. T-cell-mediated regulation of osteoclastogenesis

by signalling cross-talk between RANKL and IFN-gamma. Nature 2000;408:600-605.

39. Alnaeeli M, Penninger JM, Teng YT. Immune interactions with CD4+ T cells promote the

development of functional osteoclasts from murine CD11c+ dendritic cells. J.Immunol.

2006;177:3314-3326.

40. Lebre MC, Tak PP. Dendritic cell subsets: their roles in rheumatoid arthritis. Acta

Reumatol.Port. 2008;33:35-45.

41. Maitra R, Follenzi A, Yaghoobian A et al. Dendritic cell-mediated in vivo bone resorption.

J.Immunol. 2010;185:1485-1491.

42. Alnaeeli M, Teng YT. Dendritic cells differentiate into osteoclasts in bone marrow

microenvironment in vivo. Blood 2009;113:264-265.

43. Goldring SR. Pathogenesis of bone and cartilage destruction in rheumatoid arthritis.

Rheumatology.(Oxford) 2003;42 Suppl 2:ii11-ii16.

44. Firestein GS. Evolving concepts of rheumatoid arthritis. Nature 2003;423:356-361.

45. Holmdahl R, Malmstrom V, Burkhardt H. Autoimmune priming tissue attack and chronic

inflammation - The three stages of rheumatoid arthritis. Eur.J.Immunol. 2014

46. Eyre S, Bowes J, Diogo D et al. High-density genetic mapping identifies new susceptibility

loci for rheumatoid arthritis. Nat.Genet. 2012;44:1336-1340.

101

47. Stahl EA, Raychaudhuri S, Remmers EF et al. Genome-wide association study meta-

analysis identifies seven new rheumatoid arthritis risk loci. Nat.Genet. 2010;42:508-514.

48. Zhernakova A, Stahl EA, Trynka G et al. Meta-analysis of genome-wide association studies

in celiac disease and rheumatoid arthritis identifies fourteen non-HLA shared loci.

PLoS.Genet. 2011;7:e1002004.

49. Gregersen PK, Silver J, Winchester RJ. The shared epitope hypothesis. An approach to

understanding the molecular genetics of susceptibility to rheumatoid arthritis. Arthritis

Rheum. 1987;30:1205-1213.

50. Rioux JD, Abbas AK. Paths to understanding the genetic basis of autoimmune disease.

Nature 2005;435:584-589.

51. Karlson EW, Chang SC, Cui J et al. Gene-environment interaction between HLA-DRB1

shared epitope and heavy cigarette smoking in predicting incident rheumatoid arthritis.

Ann.Rheum.Dis. 2010;69:54-60.

52. Scally SW, Petersen J, Law SC et al. A molecular basis for the association of the HLA-

DRB1 locus, citrullination, and rheumatoid arthritis. J.Exp.Med. 2013;210:2569-2582.

53. McInnes IB, Schett G. The pathogenesis of rheumatoid arthritis. N.Engl.J.Med.

2011;365:2205-2219.

54. Schett G. Cells of the synovium in rheumatoid arthritis. Osteoclasts. Arthritis Res.Ther.

2007;9:203.

55. Strand V, Kimberly R, Isaacs JD. Biologic therapies in rheumatology: lessons learned,

future directions. Nat.Rev.Drug Discov. 2007;6:75-92.

56. Zhou M, Qin S, Chu Y et al. Immunolocalization of MMP-2 and MMP-9 in human

rheumatoid synovium. Int.J.Clin.Exp.Pathol. 2014;7:3048-3056.

57. Tran CN, Lundy SK, Fox DA. Synovial biology and T cells in rheumatoid arthritis.

Pathophysiology. 2005;12:183-189.

58. Lutzky V, Hannawi S, Thomas R. Cells of the synovium in rheumatoid arthritis. Dendritic

cells. Arthritis Res.Ther. 2007;9:219.

59. McKenna HJ, Stocking KL, Miller RE et al. Mice lacking flt3 ligand have deficient

hematopoiesis affecting hematopoietic progenitor cells, dendritic cells, and natural killer

cells. Blood 2000;95:3489-3497.

60. Page G, Lebecque S, Miossec P. Anatomic localization of immature and mature dendritic

cells in an ectopic lymphoid organ: correlation with selective chemokine expression in

rheumatoid synovium. J.Immunol. 2002;168:5333-5341.

61. Page G, Miossec P. RANK and RANKL expression as markers of dendritic cell-T cell

interactions in paired samples of rheumatoid synovium and lymph nodes. Arthritis Rheum.

2005;52:2307-2312.

102

62. Jongbloed SL, Lebre MC, Fraser AR et al. Enumeration and phenotypical analysis of

distinct dendritic cell subsets in psoriatic arthritis and rheumatoid arthritis. Arthritis

Res.Ther. 2006;8:R15.

63. Radstake TR, van Lieshout AW, van Riel PL, van den Berg WB, Adema GJ. Dendritic cells,

Fc{gamma} receptors, and Toll-like receptors: potential allies in the battle against

rheumatoid arthritis. Ann.Rheum.Dis. 2005;64:1532-1538.

64. Leung BP, Conacher M, Hunter D et al. A novel dendritic cell-induced model of erosive

inflammatory arthritis: distinct roles for dendritic cells in T cell activation and induction of

local inflammation. J.Immunol. 2002;169:7071-7077.

65. Sato K, Suematsu A, Okamoto K et al. Th17 functions as an osteoclastogenic helper T cell

subset that links T cell activation and bone destruction. J.Exp.Med. 2006;203:2673-2682.

66. Kotake S, Udagawa N, Takahashi N et al. IL-17 in synovial fluids from patients with

rheumatoid arthritis is a potent stimulator of osteoclastogenesis. J.Clin.Invest

1999;103:1345-1352.

67. Edwards JC, Szczepanski L, Szechinski J et al. Efficacy of B-cell-targeted therapy with

rituximab in patients with rheumatoid arthritis. N.Engl.J.Med. 2004;350:2572-2581.

68. Dayer JM, Bresnihan B. Targeting interleukin-1 in the treatment of rheumatoid arthritis.

Arthritis Rheum. 2002;46:574-578.

69. van den Berg WB, Joosten LA, Kollias G, van de Loo FA. Role of tumour necrosis factor

alpha in experimental arthritis: separate activity of interleukin 1beta in chronicity and

cartilage destruction. Ann.Rheum.Dis. 1999;58 Suppl 1:I40-I48.

70. Yang G, Im HJ, Wang JH. Repetitive mechanical stretching modulates IL-1beta induced

COX-2, MMP-1 expression, and PGE2 production in human patellar tendon fibroblasts.

Gene 2005;363:166-172.

71. Sack U, Kinne RW, Marx T et al. Interleukin-6 in synovial fluid is closely associated with

chronic synovitis in rheumatoid arthritis. Rheumatol.Int. 1993;13:45-51.

72. Madhok R, Crilly A, Watson J, Capell HA. Serum interleukin 6 levels in rheumatoid

arthritis: correlations with clinical and laboratory indices of disease activity.

Ann.Rheum.Dis. 1993;52:232-234.

73. Houssiau FA, Devogelaer JP, Van DJ, de Deuxchaisnes CN, Van SJ. Interleukin-6 in

synovial fluid and serum of patients with rheumatoid arthritis and other inflammatory

arthritides. Arthritis Rheum. 1988;31:784-788.

74. Nakahara H, Song J, Sugimoto M et al. Anti-interleukin-6 receptor antibody therapy reduces

vascular endothelial growth factor production in rheumatoid arthritis. Arthritis Rheum.

2003;48:1521-1529.

75. Feldmann M, Brennan FM, Maini RN. Role of cytokines in rheumatoid arthritis.

Annu.Rev.Immunol. 1996;14:397-440.

103

76. Robinson E, Keystone EC, Schall TJ, Gillett N, Fish EN. Chemokine expression in

rheumatoid arthritis (RA): evidence of RANTES and macrophage inflammatory protein

(MIP)-1 beta production by synovial T cells. Clin.Exp.Immunol. 1995;101:398-407.

77. Loetscher P, Seitz M, Clark-Lewis I, Baggiolini M, Moser B. Monocyte chemotactic

proteins MCP-1, MCP-2, and MCP-3 are major attractants for human CD4+ and CD8+ T

lymphocytes. FASEB J. 1994;8:1055-1060.

78. Koch AE, Kunkel SL, Harlow LA et al. Enhanced production of monocyte chemoattractant

protein-1 in rheumatoid arthritis. J.Clin.Invest 1992;90:772-779.

79. Rathanaswami P, Hachicha M, Sadick M, Schall TJ, McColl SR. Expression of the cytokine

RANTES in human rheumatoid synovial fibroblasts. Differential regulation of RANTES

and interleukin-8 genes by inflammatory cytokines. J.Biol.Chem. 1993;268:5834-5839.

80. Iwamoto T, Okamoto H, Toyama Y, Momohara S. Molecular aspects of rheumatoid

arthritis: chemokines in the joints of patients. FEBS J. 2008;275:4448-4455.

81. Koch AE, Kunkel SL, Harlow LA et al. Macrophage inflammatory protein-1 alpha. A novel

chemotactic cytokine for macrophages in rheumatoid arthritis. J.Clin.Invest 1994;93:921-

928.

82. Toussirot E, Wendling D. The use of TNF-alpha blocking agents in rheumatoid arthritis: an

update. Expert.Opin.Pharmacother. 2007;8:2089-2107.

83. Choi Y, Arron JR, Townsend MJ. Promising bone-related therapeutic targets for rheumatoid

arthritis. Nat.Rev.Rheumatol. 2009;5:543-548.

84. Nell VP, Machold KP, Eberl G et al. Benefit of very early referral and very early therapy

with disease-modifying anti-rheumatic drugs in patients with early rheumatoid arthritis.

Rheumatology.(Oxford) 2004;43:906-914.

85. Maini RN, Taylor PC. Anti-cytokine therapy for rheumatoid arthritis. Annu.Rev.Med.

2000;51:207-229.

86. Lipsky PE, van der Heijde DM, St Clair EW et al. Infliximab and methotrexate in the

treatment of rheumatoid arthritis. Anti-Tumor Necrosis Factor Trial in Rheumatoid Arthritis

with Concomitant Therapy Study Group. N.Engl.J.Med. 2000;343:1594-1602.

87. Redlich K, Hayer S, Ricci R et al. Osteoclasts are essential for TNF-alpha-mediated joint

destruction. J.Clin.Invest 2002;110:1419-1427.

88. Ehrenstein MR, Evans JG, Singh A et al. Compromised function of regulatory T cells in

rheumatoid arthritis and reversal by anti-TNFalpha therapy. J.Exp.Med. 2004;200:277-285.

89. Feldmann M, Brennan FM, Foxwell BM et al. Anti-TNF therapy: where have we got to in

2005? J.Autoimmun. 2005;25 Suppl:26-28.

90. Kelly V, Genovese M. Novel small molecule therapeutics in rheumatoid arthritis.

Rheumatology.(Oxford) 2013;52:1155-1162.

104

91. van der Heijde D, Tanaka Y, Fleischmann R et al. Tofacitinib (CP-690,550) in patients with

rheumatoid arthritis receiving methotrexate: twelve-month data from a twenty-four-month

phase III randomized radiographic study. Arthritis Rheum. 2013;65:559-570.

92. McClung MR, Lewiecki EM, Cohen SB et al. Denosumab in postmenopausal women with

low bone mineral density. N.Engl.J.Med. 2006;354:821-831.

93. McClung MR. Inhibition of RANKL as a treatment for osteoporosis: preclinical and early

clinical studies. Curr.Osteoporos.Rep. 2006;4:28-33.

94. Cohen SB, Dore RK, Lane NE et al. Denosumab treatment effects on structural damage,

bone mineral density, and bone turnover in rheumatoid arthritis: a twelve-month,

multicenter, randomized, double-blind, placebo-controlled, phase II clinical trial. Arthritis

Rheum. 2008;58:1299-1309.

95. Aumailley M, Gayraud B. Structure and biological activity of the extracellular matrix.

J.Mol.Med.(Berl) 1998;76:253-265.

96. Myllyharju J, Kivirikko KI. Collagens, modifying enzymes and their mutations in humans,

flies and worms. Trends Genet. 2004;20:33-43.

97. Karsdal MA, Nielsen MJ, Sand JM et al. Extracellular matrix remodeling: the common

denominator in connective tissue diseases. Possibilities for evaluation and current

understanding of the matrix as more than a passive architecture, but a key player in tissue

failure. Assay.Drug Dev.Technol. 2013;11:70-92.

98. Gelse K, Poschl E, Aigner T. Collagens--structure, function, and biosynthesis. Adv.Drug

Deliv.Rev. 2003;55:1531-1546.

99. Leitinger B. Transmembrane collagen receptors. Annu.Rev.Cell Dev.Biol. 2011;27:265-290.

100. Pacifici R, Basilico C, Roman J et al. Collagen-induced release of interleukin 1 from human

blood mononuclear cells. Potentiation by fibronectin binding to the alpha 5 beta 1 integrin.

J.Clin.Invest 1992;89:61-67.

101. Takayanagi H. Mechanistic insight into osteoclast differentiation in osteoimmunology.

J.Mol.Med.(Berl) 2005;83:170-179.

102. Yasuda T. Cartilage destruction by matrix degradation products. Mod.Rheumatol.

2006;16:197-205.

103. Sophia Fox AJ, Bedi A, Rodeo SA. The basic science of articular cartilage: structure,

composition, and function. Sports Health 2009;1:461-468.

104. Goldring MB, Marcu KB. Cartilage homeostasis in health and rheumatic diseases. Arthritis

Res.Ther. 2009;11:224.

105. Brodsky B, Persikov AV. Molecular structure of the collagen triple helix. Adv.Protein

Chem. 2005;70:301-339.

105

106. Shoulders MD, Raines RT. Collagen structure and stability. Annu.Rev.Biochem.

2009;78:929-958.

107. Ricard-Blum S. The collagen family. Cold Spring Harb.Perspect.Biol. 2011;3:a004978.

108. Lu P, Takai K, Weaver VM, Werb Z. Extracellular matrix degradation and remodeling in

development and disease. Cold Spring Harb.Perspect.Biol. 2011;3:

109. Cawston TE, Young DA. Proteinases involved in matrix turnover during cartilage and bone

breakdown. Cell Tissue Res. 2010;339:221-235.

110. Parks WC, Wilson CL, Lopez-Boado YS. Matrix metalloproteinases as modulators of

inflammation and innate immunity. Nat.Rev.Immunol. 2004;4:617-629.

111. Nagase H, Visse R, Murphy G. Structure and function of matrix metalloproteinases and

TIMPs. Cardiovasc.Res. 2006;69:562-573.

112. Burrage PS, Mix KS, Brinckerhoff CE. Matrix metalloproteinases: role in arthritis. Front

Biosci. 2006;11:529-543.

113. Arend WP, Dayer JM. Inhibition of the production and effects of interleukin-1 and tumor

necrosis factor alpha in rheumatoid arthritis. Arthritis Rheum. 1995;38:151-160.

114. Yasuda T. Cartilage destruction by matrix degradation products. Mod.Rheumatol.

2006;16:197-205.

115. Davis GE, Pintar Allen KA, Salazar R, Maxwell SA. Matrix metalloproteinase-1 and -9

activation by plasmin regulates a novel endothelial cell-mediated mechanism of collagen gel

contraction and capillary tube regression in three-dimensional collagen matrices. J.Cell Sci.

2001;114:917-930.

116. Ramos-DeSimone N, Hahn-Dantona E, Sipley J et al. Activation of matrix

metalloproteinase-9 (MMP-9) via a converging plasmin/stromelysin-1 cascade enhances

tumor cell invasion. J.Biol.Chem. 1999;274:13066-13076.

117. Brew K, Dinakarpandian D, Nagase H. Tissue inhibitors of metalloproteinases: evolution,

structure and function. Biochim.Biophys.Acta 2000;1477:267-283.

118. Smith HW, Marshall CJ. Regulation of cell signalling by uPAR. Nat.Rev.Mol.Cell Biol.

2010;11:23-36.

119. Green KA, Lund LR. ECM degrading proteases and tissue remodelling in the mammary

gland. Bioessays 2005;27:894-903.

120. Carmeliet P, Moons L, Lijnen R et al. Urokinase-generated plasmin activates matrix

metalloproteinases during aneurysm formation. Nat.Genet. 1997;17:439-444.

121. Gong Y, Hart E, Shchurin A, Hoover-Plow J. Inflammatory macrophage migration requires

MMP-9 activation by plasminogen in mice. J.Clin.Invest 2008;118:3012-3024.

106

122. Madsen DH, Leonard D, Masedunskas A et al. M2-like macrophages are responsible for

collagen degradation through a mannose receptor-mediated pathway. J.Cell Biol.

2013;202:951-966.

123. Rao JS. Molecular mechanisms of glioma invasiveness: the role of proteases.

Nat.Rev.Cancer 2003;3:489-501.

124. Hynes RO. Integrins: bidirectional, allosteric signaling machines. Cell 2002;110:673-687.

125. Vogel WF, Abdulhussein R, Ford CE. Sensing extracellular matrix: an update on discoidin

domain receptor function. Cell Signal. 2006;18:1108-1116.

126. Ferri N, Carragher NO, Raines EW. Role of discoidin domain receptors 1 and 2 in human

smooth muscle cell-mediated collagen remodeling: potential implications in atherosclerosis

and lymphangioleiomyomatosis. Am.J.Pathol. 2004;164:1575-1585.

127. Olaso E, Labrador JP, Wang L et al. Discoidin domain receptor 2 regulates fibroblast

proliferation and migration through the extracellular matrix in association with

transcriptional activation of matrix metalloproteinase-2. J.Biol.Chem. 2002;277:3606-3613.

128. Brondijk TH, de RT, Ballering J et al. Crystal structure and collagen-binding site of immune

inhibitory receptor LAIR-1: unexpected implications for collagen binding by platelet

receptor GPVI. Blood 2010;115:1364-1373.

129. Farndale RW, Slatter DA, Siljander PR, Jarvis GE. Platelet receptor recognition and cross-

talk in collagen-induced activation of platelets. J.Thromb.Haemost. 2007;5 Suppl 1:220-

229.

130. Meyaard L. LAIR and collagens in immune regulation. Immunol.Lett. 2010;128:26-28.

131. Xu H, Bihan D, Chang F et al. Discoidin domain receptors promote alpha1beta1- and

alpha2beta1-integrin mediated cell adhesion to collagen by enhancing integrin activation.

PLoS.One. 2012;7:e52209.

132. Jeng KC, Liu MT, Lan JL et al. Collagen induces cytokine production by synovial fluid

mononuclear cells in rheumatoid arthritis. Immunol.Lett. 1995;45:13-17.

133. Dietrich J, Nakajima H, Colonna M. Human inhibitory and activating Ig-like receptors

which modulate the function of myeloid cells. Microbes.Infect. 2000;2:323-329.

134. Merck E, Gaillard C, Scuiller M et al. Ligation of the FcR gamma chain-associated human

osteoclast-associated receptor enhances the proinflammatory responses of human monocytes

and neutrophils. J.Immunol. 2006;176:3149-3156.

135. Ishikawa S, Arase N, Suenaga T et al. Involvement of FcRgamma in signal transduction of

osteoclast-associated receptor (OSCAR). Int.Immunol. 2004;16:1019-1025.

136. Merck E, Gaillard C, Gorman DM et al. OSCAR is an FcRgamma-associated receptor that

is expressed by myeloid cells and is involved in antigen presentation and activation of

human dendritic cells. Blood 2004;104:1386-1395.

107

137. Barrow AD, Raynal N, Andersen TL et al. OSCAR is a collagen receptor that costimulates

osteoclastogenesis in DAP12-deficient humans and mice. J.Clin.Invest 2011;121:3505-

3516.

138. Zhao S, Guo YY, Ding N, Yang LL, Zhang N. Changes in serum levels of soluble

osteoclast-associated receptor in human rheumatoid arthritis. Chin Med.J.(Engl.)

2011;124:3058-3060.

139. Crotti TN, Dharmapatni AA, Alias E et al. The immunoreceptor tyrosine-based activation

motif (ITAM) -related factors are increased in synovial tissue and vasculature of rheumatoid

arthritic joints. Arthritis Res.Ther. 2012;14:R245.

140. Ndongo-Thiam N, de SG, Kastrup J, Miossec P. Levels of soluble osteoclast-associated

receptor (sOSCAR) in rheumatoid arthritis: link to disease severity and cardiovascular risk.

Ann.Rheum.Dis. 2014;73:1276-1277.

141. Herman S, Muller RB, Kronke G et al. Induction of osteoclast-associated receptor, a key

osteoclast costimulation molecule, in rheumatoid arthritis. Arthritis Rheum. 2008;58:3041-

3050.

142. Schultz HS, Nitze LM, Zeuthen LH et al. Collagen Induces Maturation of Human

Monocyte-Derived Dendritic Cells by Signaling through Osteoclast-Associated Receptor.

J.Immunol. 2015

143. Bonecchi R, Bianchi G, Bordignon PP et al. Differential expression of chemokine receptors

and chemotactic responsiveness of type 1 T helper cells (Th1s) and Th2s. J.Exp.Med.

1998;187:129-134.

144. Iellem A, Mariani M, Lang R et al. Unique chemotactic response profile and specific

expression of chemokine receptors CCR4 and CCR8 by CD4(+)CD25(+) regulatory T cells.

J.Exp.Med. 2001;194:847-853.

145. So H, Rho J, Jeong D et al. Microphthalmia transcription factor and PU.1 synergistically

induce the leukocyte receptor osteoclast-associated receptor gene expression. J.Biol.Chem.

2003;278:24209-24216.

146. Kim GS, Koh JM, Chang JS et al. Association of the OSCAR promoter polymorphism with

BMD in postmenopausal women. J.Bone Miner.Res. 2005;20:1342-1348.

147. Goettsch C, Rauner M, Sinningen K et al. The osteoclast-associated receptor (OSCAR) is a

novel receptor regulated by oxidized low-density lipoprotein in human endothelial cells.

Endocrinology 2011;152:4915-4926.

148. Goettsch C, Kliemt S, Sinningen K et al. Quantitative proteomics reveals novel functions of

osteoclast-associated receptor in STAT signaling and cell adhesion in human endothelial

cells. J.Mol.Cell Cardiol. 2012;53:829-837.

149. Sinningen K, Rauner M, Goettsch C et al. Monocytic expression of osteoclast-associated

receptor (OSCAR) is induced in atherosclerotic mice and regulated by oxidized low-density

lipoprotein in vitro. Biochem.Biophys.Res.Commun. 2013;437:314-318.

108

150. van den Berg WB. Animal models of arthritis. What have we learned? J.Rheumatol.Suppl

2005;72:7-9.

151. Liote F, Boval-Boizard B, Weill D, Kuntz D, Wautier JL. Blood monocyte activation in

rheumatoid arthritis: increased monocyte adhesiveness, integrin expression, and cytokine

release. Clin.Exp.Immunol. 1996;106:13-19.

152. Torsteinsdottir I, Arvidson NG, Hallgren R, Hakansson L. Monocyte activation in

rheumatoid arthritis (RA): increased integrin, Fc gamma and complement receptor

expression and the effect of glucocorticoids. Clin.Exp.Immunol. 1999;115:554-560.

153. Vieira-Sousa E, Gerlag DM, Tak PP. Synovial tissue response to treatment in rheumatoid

arthritis. Open.Rheumatol.J. 2011;5:115-122.

154. Mulherin D, Fitzgerald O, Bresnihan B. Synovial tissue macrophage populations and

articular damage in rheumatoid arthritis. Arthritis Rheum. 1996;39:115-124.

155. Haringman JJ, Gerlag DM, Zwinderman AH et al. Synovial tissue macrophages: a sensitive

biomarker for response to treatment in patients with rheumatoid arthritis. Ann.Rheum.Dis.

2005;64:834-838.

156. Davignon JL, Hayder M, Baron M et al. Targeting monocytes/macrophages in the treatment

of rheumatoid arthritis. Rheumatology.(Oxford) 2013;52:590-598.

157. Mosser DM, Edwards JP. Exploring the full spectrum of macrophage activation.

Nat.Rev.Immunol. 2008;8:958-969.

158. Cheung HS, Ryan LM, Kozin F, McCarty DJ. Identification of collagen subtypes in

synovial fluid sediments from arthritic patients. Am.J.Med. 1980;68:73-79.

159. Wotton SF, Dieppe PA, Duance VC. Type IX collagen immunoreactive peptides in synovial

fluids from arthritis patients. Rheumatology.(Oxford) 1999;38:338-345.

160. Lohmander LS, Atley LM, Pietka TA, Eyre DR. The release of crosslinked peptides from

type II collagen into human synovial fluid is increased soon after joint injury and in

osteoarthritis. Arthritis Rheum. 2003;48:3130-3139.

161. Canfield SM, Morrison SL. The binding affinity of human IgG for its high affinity Fc

receptor is determined by multiple amino acids in the CH2 domain and is modulated by the

hinge region. J.Exp.Med. 1991;173:1483-1491.

162. Bleharski JR, Kiessler V, Buonsanti C et al. A role for triggering receptor expressed on

myeloid cells-1 in host defense during the early-induced and adaptive phases of the immune

response. J.Immunol. 2003;170:3812-3818.

163. Menzies-Gow A, Ying S, Sabroe I et al. Eotaxin (CCL11) and eotaxin-2 (CCL24) induce

recruitment of eosinophils, basophils, neutrophils, and macrophages as well as features of

early- and late-phase allergic reactions following cutaneous injection in human atopic and

nonatopic volunteers. J.Immunol. 2002;169:2712-2718.

109

164. Pacifici R, Carano A, Santoro SA et al. Bone matrix constituents stimulate interleukin-1

release from human blood mononuclear cells. J.Clin.Invest 1991;87:221-228.

165. Hynes RO. The extracellular matrix: not just pretty fibrils. Science 2009;326:1216-1219.

166. Daley WP, Peters SB, Larsen M. Extracellular matrix dynamics in development and

regenerative medicine. J.Cell Sci. 2008;121:255-264.

167. Karsdal MA, Woodworth T, Henriksen K et al. Biochemical markers of ongoing joint

damage in rheumatoid arthritis--current and future applications, limitations and

opportunities. Arthritis Res.Ther. 2011;13:215.

168. Neve A, Corrado A, Cantatore FP. TNF-related apoptosis-inducing ligand (TRAIL) in

rheumatoid arthritis: what's new? Clin.Exp.Med. 2014;14:115-120.

169. Baldwin HM, Ito-Ihara T, Isaacs JD, Hilkens CM. Tumour necrosis factor alpha blockade

impairs dendritic cell survival and function in rheumatoid arthritis. Ann.Rheum.Dis.

2010;69:1200-1207.

170. Lo SZ, Steer JH, Joyce DA. Tumor necrosis factor-alpha promotes survival in methotrexate-

exposed macrophages by an NF-kappaB-dependent pathway. Arthritis Res.Ther.

2011;13:R24.

171. Flad HD, Grage-Griebenow E, Petersen F et al. The role of cytokines in monocyte

apoptosis. Pathobiology 1999;67:291-293.

172. Hounoki H, Sugiyama E, Mohamed SG et al. Activation of peroxisome proliferator-

activated receptor gamma inhibits TNF-alpha-mediated osteoclast differentiation in human

peripheral monocytes in part via suppression of monocyte chemoattractant protein-1

expression. Bone 2008;42:765-774.

173. Szekanecz Z, Koch AE. Macrophages and their products in rheumatoid arthritis.

Curr.Opin.Rheumatol. 2007;19:289-295.

174. Fahey TJ, III, Tracey KJ, Tekamp-Olson P et al. Macrophage inflammatory protein 1

modulates macrophage function. J.Immunol. 1992;148:2764-2769.

175. Narni-Mancinelli E, Soudja SM, Crozat K et al. Inflammatory monocytes and neutrophils

are licensed to kill during memory responses in vivo. PLoS.Pathog. 2011;7:e1002457.

176. Fantuzzi L, Borghi P, Ciolli V et al. Loss of CCR2 expression and functional response to

monocyte chemotactic protein (MCP-1) during the differentiation of human monocytes: role

of secreted MCP-1 in the regulation of the chemotactic response. Blood 1999;94:875-883.

177. Katschke KJ, Jr., Rottman JB, Ruth JH et al. Differential expression of chemokine receptors

on peripheral blood, synovial fluid, and synovial tissue monocytes/macrophages in

rheumatoid arthritis. Arthritis Rheum. 2001;44:1022-1032.

178. Phillips RJ, Lutz M, Premack B. Differential signaling mechanisms regulate expression of

CC chemokine receptor-2 during monocyte maturation. J.Inflamm.(Lond) 2005;2:14.

110

179. Bouchon A, Dietrich J, Colonna M. Cutting edge: inflammatory responses can be triggered

by TREM-1, a novel receptor expressed on neutrophils and monocytes. J.Immunol.

2000;164:4991-4995.

180. Knapp S, Gibot S, de VA et al. Cutting edge: expression patterns of surface and soluble

triggering receptor expressed on myeloid cells-1 in human endotoxemia. J.Immunol.

2004;173:7131-7134.

181. Sharif O, Bolshakov VN, Raines S, Newham P, Perkins ND. Transcriptional profiling of the

LPS induced NF-kappaB response in macrophages. BMC.Immunol. 2007;8:1.

182. Netea MG, Azam T, Ferwerda G et al. Triggering receptor expressed on myeloid cells-1

(TREM-1) amplifies the signals induced by the NACHT-LRR (NLR) pattern recognition

receptors. J.Leukoc.Biol. 2006;80:1454-1461.

183. Perkins ND. Integrating cell-signalling pathways with NF-kappaB and IKK function.

Nat.Rev.Mol.Cell Biol. 2007;8:49-62.

184. Kuai J, Gregory B, Hill A et al. TREM-1 expression is increased in the synovium of

rheumatoid arthritis patients and induces the expression of pro-inflammatory cytokines.

Rheumatology.(Oxford) 2009;48:1352-1358.

185. Murakami Y, Akahoshi T, Aoki N et al. Intervention of an inflammation amplifier,

triggering receptor expressed on myeloid cells 1, for treatment of autoimmune arthritis.

Arthritis Rheum. 2009;60:1615-1623.

186. Sedlacek R, Mauch S, Kolb B et al. Matrix metalloproteinase MMP-19 (RASI-1) is

expressed on the surface of activated peripheral blood mononuclear cells and is detected as

an autoantigen in rheumatoid arthritis. Immunobiology 1998;198:408-423.

187. Kolb C, Mauch S, Peter HH, Krawinkel U, Sedlacek R. The matrix metalloproteinase RASI-

1 is expressed in synovial blood vessels of a rheumatoid arthritis patient. Immunol.Lett.

1997;57:83-88.

188. Stracke JO, Hutton M, Stewart M et al. Biochemical characterization of the catalytic domain

of human matrix metalloproteinase 19. Evidence for a role as a potent basement membrane

degrading enzyme. J.Biol.Chem. 2000;275:14809-14816.

189. Stracke JO, Fosang AJ, Last K et al. Matrix metalloproteinases 19 and 20 cleave aggrecan

and cartilage oligomeric matrix protein (COMP). FEBS Lett. 2000;478:52-56.

190. Hakala M, Risteli L, Manelius J, Nieminen P, Risteli J. Increased type I collagen

degradation correlates with disease severity in rheumatoid arthritis. Ann.Rheum.Dis.

1993;52:866-869.

191. Hakala M, Aman S, Luukkainen R et al. Application of markers of collagen metabolism in

serum and synovial fluid for assessment of disease process in patients with rheumatoid

arthritis. Ann.Rheum.Dis. 1995;54:886-890.

111

192. Tak PP, Breedveld FC. Analysis of serial synovial biopsies as a screening method for

predicting the effects of therapeutic interventions. J.Clin.Rheumatol. 1997;3:186.

193. Tak PP, Smeets TJ, Daha MR et al. Analysis of the synovial cell infiltrate in early

rheumatoid synovial tissue in relation to local disease activity. Arthritis Rheum.

1997;40:217-225.

194. Hamilton JA, Tak PP. The dynamics of macrophage lineage populations in inflammatory

and autoimmune diseases. Arthritis Rheum. 2009;60:1210-1221.

195. Schneider C, Newman RA, Sutherland DR, Asser U, Greaves MF. A one-step purification

of membrane proteins using a high efficiency immunomatrix. J.Biol.Chem.

1982;257:10766-10769.

196. Herenius MM, Thurlings RM, Wijbrandts CA et al. Monocyte migration to the synovium in

rheumatoid arthritis patients treated with adalimumab. Ann.Rheum.Dis. 2011;70:1160-1162.

197. Wijbrandts CA, Remans PH, Klarenbeek PL et al. Analysis of apoptosis in peripheral blood

and synovial tissue very early after initiation of infliximab treatment in rheumatoid arthritis

patients. Arthritis Rheum. 2008;58:3330-3339.

198. Verbrugge A, Meyaard L. Signaling by ITIM-Bearing Receptors. Current Immunology

Reviews 2015;1:201-212.

199. Lebbink RJ, de RT, Adelmeijer J et al. Collagens are functional, high affinity ligands for the

inhibitory immune receptor LAIR-1. J.Exp.Med. 2006;203:1419-1425.

200. Lebbink RJ, Raynal N, de RT et al. Identification of multiple potent binding sites for human

leukocyte associated Ig-like receptor LAIR on collagens II and III. Matrix Biol.

2009;28:202-210.

201. Fleetwood AJ, Achuthan A, Schultz H et al. Urokinase plasminogen activator is a central

regulator of macrophage three-dimensional invasion, matrix degradation, and adhesion.

J.Immunol. 2014;192:3540-3547.

202. Shapiro SD, Kobayashi DK, Pentland AP, Welgus HG. Induction of macrophage

metalloproteinases by extracellular matrix. Evidence for enzyme- and substrate-specific

responses involving prostaglandin-dependent mechanisms. J.Biol.Chem. 1993;268:8170-

8175.

203. Lepidi S, Kenagy RD, Raines EW et al. MMP9 production by human monocyte-derived

macrophages is decreased on polymerized type I collagen. J.Vasc.Surg. 2001;34:1111-1118.

204. Franz S, Allenstein F, Kajahn J et al. Artificial extracellular matrices composed of collagen I

and high-sulfated hyaluronan promote phenotypic and functional modulation of human pro-

inflammatory M1 macrophages. Acta Biomater. 2013;9:5621-5629.

205. Nemeth K, Schoppet M, Al-Fakhri N et al. The role of osteoclast-associated receptor in

osteoimmunology. J.Immunol. 2011;186:13-18.

112

206. Trentham DE, Kammer GM, McCune WJ, David JR. Autoimmunity to collagen: a shared

feature of psoriatic and rheumatoid arthritis. Arthritis Rheum. 1981;24:1363-1369.

207. Stuart JM, Huffstutter EH, Townes AS, Kang AH. Incidence and specificity of antibodies to

types I, II, III, IV, and V collagen in rheumatoid arthritis and other rheumatic diseases as

measured by 125I-radioimmunoassay. Arthritis Rheum. 1983;26:832-840.

208. Dalod M, Chelbi R, Malissen B, Lawrence T. Dendritic cell maturation: functional

specialization through signaling specificity and transcriptional programming. EMBO J.

2014;33:1104-1116.

209. Shortman K, Naik SH. Steady-state and inflammatory dendritic-cell development.

Nat.Rev.Immunol. 2007;7:19-30.

210. Swirski FK, Nahrendorf M, Etzrodt M et al. Identification of splenic reservoir monocytes

and their deployment to inflammatory sites. Science 2009;325:612-616.

211. Sengupta TK, Chen A, Zhong Z, Darnell JE, Jr., Ivashkiv LB. Activation of monocyte

effector genes and STAT family transcription factors by inflammatory synovial fluid is

independent of interferon gamma. J.Exp.Med. 1995;181:1015-1025.

212. Sallusto F, Lanzavecchia A, Mackay CR. Chemokines and chemokine receptors in T-cell

priming and Th1/Th2-mediated responses. Immunol.Today 1998;19:568-574.

213. Baggiolini M. Chemotactic and inflammatory cytokines--CXC and CC proteins.

Adv.Exp.Med.Biol. 1993;351:1-11.

214. Mukaida N. Pathophysiological roles of interleukin-8/CXCL8 in pulmonary diseases.

Am.J.Physiol Lung Cell Mol.Physiol 2003;284:L566-L577.

215. Ludewig B, Graf D, Gelderblom HR et al. Spontaneous apoptosis of dendritic cells is

efficiently inhibited by TRAP (CD40-ligand) and TNF-alpha, but strongly enhanced by

interleukin-10. Eur.J.Immunol. 1995;25:1943-1950.

216. Slobodin G, Kessel A, Peri R et al. Etanercept impairs maturation of human monocyte-

derived dendritic cells by inhibiting the autocrine TNFalpha-mediated signaling.

Inflammation 2009;32:146-150.

217. Sarkar S, Fox DA. Dendritic cells in rheumatoid arthritis. Front Biosci. 2005;10:656-665.

218. Turner NJ, Pezzone D, Badylak SF. Regional Variations in the Histology of Porcine Skin.

Tissue Eng Part C.Methods 2014

219. Watson RE, Gibbs NK, Griffiths CE, Sherratt MJ. Damage to skin extracellular matrix

induced by UV exposure. Antioxid.Redox.Signal. 2014;21:1063-1077.

220. Agarwal P, Zwolanek D, Keene DR et al. Collagen XII and XIV, new partners of cartilage

oligomeric matrix protein in the skin extracellular matrix suprastructure. J.Biol.Chem.

2012;287:22549-22559.

113

221. Seymour HE, Worsley A, Smith JM, Thomas SH. Anti-TNF agents for rheumatoid arthritis.

Br.J.Clin.Pharmacol. 2001;51:201-208.

222. Curtis JR, Singh JA. Use of biologics in rheumatoid arthritis: current and emerging

paradigms of care. Clin.Ther. 2011;33:679-707.

223. Lusis AJ. Atherosclerosis. Nature 2000;407:233-241.

224. Niessner A, Weyand CM. Dendritic cells in atherosclerotic disease. Clin.Immunol.

2010;134:25-32.

225. Bobryshev YV, Lord RS. S-100 positive cells in human arterial intima and in atherosclerotic

lesions. Cardiovasc.Res. 1995;29:689-696.

226. Yilmaz A, Lochno M, Traeg F et al. Emergence of dendritic cells in rupture-prone regions

of vulnerable carotid plaques. Atherosclerosis 2004;176:101-110.

114

8. Appendix

8.1 Appendix I

Supplemental figures and tables for:

Heidi S. Schultz, Louise M. Nitze,, Louise H. Zeuthen, Pernille Keller, Albrecht Gruhler, Jesper

Pass, Jianhe Chen, Li Guo, Andrew J. Fleetwood, John A. Hamilton, Martin W. Berchtold and

Svetlana Panina. (2015) Collagen Induces Maturation of Human Monocyte-Derived Dendritic

Cells by Signaling through Osteoclast-Associated. J. Immunology. 194 (epub, doi 10.4049/

jimmunol.1402800)

Figure S1. Identification of OSCAR ligand. (A) Binding of Alexa488-labelled OSCAR-ECD to the surface of live

cells was analysed by flow cytometry. A murine MC3T3-E1 pre-osteoblast cell line expressed OSCAR-binding

moieties on the cell surface and therefore was selected for purification of the OSCAR ligand. Open histograms –

fluorescence in the presence of the Alexa488-conjugated protein, black histograms - untreated cells; (B). Proteins

extracted from the membranes of the MC3T3-E1 cells were incubated with either Protein G/OSCAR-Fc affinity matrix

or control Protein G/Fc matrix. The bound proteins were eluted and analyzed by SDS-PAGE. The proteins which were

differentially bound to the OSCAR-Fc and not to the Protein G/Fc matrix were further analyzed. The major diffuse

protein band with an approx. electrophoretic mobility of 170-200 kD was subjected to in-gel tryptic digest and the

peptides analyzed by LCMSMS. Peak lists were generated and searched with MASCOT against the mammalian part of

the Swissprot database. The top hit was mouse alpha 1 type I Collagen (accession number P11087) with a MASCOT

score of 2241 and sequence coverage of 49%. In addition, mouse alpha 1 Collagen type III (accession number P08121)

was identified with a lower score of 471 and a sequence coverage of 15%. SDS-PAGE analysis of proteins from

MC3T3-E1 bound to the OSCAR-Fc (lane 1) and to the Fc control (lane 2) affinity matrix. The arrow is indicating a

band which was identified by LC-MS MS as Collagen alpha 1 type I and Collagen alpha 1 type III chains; (C).

Interaction of OSCAR-ECD with native human Collagens was further verified using surface plasmon resonance

technique (SPR). Human ColI-IV proteins were immobilized to a sensor chip followed by injection of OSCAR-ECD.

Binding of OSCAR-ECD was observed to ColI and II, but not to ColIII and IV. Thus, ColI and II appear to be natural

ligands for OSCAR.

115

Figure S2. SPR analysis of OSCAR binding to ColI in the presence of anti-OSCAR mAb. A panel of mAb against

OSCAR was generated and tested for their ability to block the interaction between OSCAR-ECD and ColI. For this

purpose, ColI was immobilized on a SPR sensor chip. Binding of a constant amount of OSCAR-ECD to ColI was

measured in the presence of increasing concentrations of anti-OSCAR mAb. In cases where the antibody did not inhibit

the interaction between OSCAR and Collagen, the SPR signal increased with increasing amounts of antibody, due to the

higher mass of the OSCAR-ECD/mAb complex. In contrast, several mAb prevented the binding of OSCAR to Collagen

thereby leading to a concentration dependent decrease of the SPR signal. The mAb from these two groups were used in

this study as control and a function-blocking mAb, respectively.

Microarray data for Affymetrix Human Genome U133 Plus PM Array can be found at:

http://www.ebi.ac.uk/arrayexpress/experiments/E-MTAB-2904/

116

8.2 Appendix II

Supplemental figures and tables for:

Heidi S. Schultz, Li Guo, Pernille Keller, Andrew J. Fleetwood, Wei Guo, John A. Hamilton, Olle

Björkdahl, Martin W. Berchtold and Svetlana Panina. OSCAR-collagen signaling mediates

activation of monocytes. To be submitted to J. Immunology March, 2015.

Figure S1. TREM-1 is the top biological upstream regulator of the genes regulated by OSCAR-ColII signaling.

Bright orange fillings indicate predicted activation and pale orange indicate less predicted activation. Pale blue fillings

indicate less predicted inhibition. The lines indicate predicted relationships: orange, leads to activation; blue, leads to

inhibition; Grey, effect not predicted and yellow, findings inconsistent with state of downstream molecule.

Microarray data for Affymetrix Human Genome U133 Plus PM Array can be found at:

http://www.ebi.ac.uk/arrayexpress/experiments/E-MTAB-3281/

117

8.3 Co-author contributions

120

D E P A R T M E N T O F C E L L A N D D E V E L O P M E N T A L B I O L O G Y

F A C U L T Y O F S C I E N C E

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

P H D T H E S I S 2 0 1 5

H E I D I S C H I Ø L E R S C H U L T Z