receptor-mediated apoptosis in b cells · eeva, jonna. receptor-mediated apoptosis in b cells. from...

92
Doctoral dissertation To be presented by permission of the Faculty of Medicine of the University of Kuopio for public examination in Auditorium L22, Snellmania building, University of Kuopio, on Friday 13 th November 2009, at 12 noon Institute of Clinical Medicine Department of Clinical Microbiology University of Kuopio JONNA EEVA Receptor-Mediated Apoptosis in B Cells From the Regulation of B Cell Survival to Potential Novel Approaches for Lymphoma Therapy JOKA KUOPIO 2009 KUOPION YLIOPISTON JULKAISUJA D. LÄÄKETIEDE 463 KUOPIO UNIVERSITY PUBLICATIONS D. MEDICAL SCIENCES 463

Upload: nguyenkhanh

Post on 02-Dec-2018

230 views

Category:

Documents


0 download

TRANSCRIPT

Doctoral dissertation

To be presented by permission of the Faculty of Medicine of the University of Kuopio

for public examination in Auditorium L22, Snellmania building, University of Kuopio,

on Friday 13th November 2009, at 12 noon

Institute of Clinical MedicineDepartment of Clinical Microbiology

University of Kuopio

JONNA EEVA

Receptor-Mediated Apoptosis in B Cells

From the Regulation of B Cell Survival to Potential Novel Approaches for Lymphoma Therapy

JOKAKUOPIO 2009

KUOPION YLIOPISTON JULKAISUJA D. LÄÄKETIEDE 463KUOPIO UNIVERSITY PUBLICATIONS D. MEDICAL SCIENCES 463

Distributor : Kuopio University Library P.O. Box 1627 FI-70211 KUOPIO FINLAND Tel. +358 40 355 3430 Fax +358 17 163 410 www.uku.fi/kirjasto/julkaisutoiminta/julkmyyn.shtml

Series Editors: Professor Raimo Sulkava, M.D., Ph.D. School of Public Health and Clinical Nutrition Professor Markku Tammi, M.D., Ph.D. Institute of Biomedicine, Department of Anatomy

Author´s address: Institute of Clinical Medicine Department of Clinical Microbiology University of Kuopio P.O. Box 1627 FI-70211 KUOPIO FINLAND E-mail : eeva@hytti .uku.fi Supervisors: Professor Jukka Pelkonen, M.D., Ph.D. Institute of Clinical Medicine Department of Clinical Microbiology University of Kuopio

Mine Eray, M.D., Ph.D. School of Medicine, University of Tampere Department of Pathology, Tampere University Hospital

Reviewers: Professor Marko Salmi, M.D., Ph.D. MediCity Research Laboratory University of Turku

Professor Mauno Vihinen, Ph.D. Institute of Medical Technology University of Tampere

Opponent: Professor Seppo Meri, M.D., Ph.D. Haartman Institute University of Helsinki

ISBN 978-951-27-1363-9ISBN 978-951-27-1380-6 (PDF)ISSN 1235-0303

KopijyväKuopio 2009Finland

Eeva, Jonna. Receptor-mediated apoptosis in B cells. From the regulation of B cell survival to

potential novel approaches for lymphoma therapy. Kuopio University Publications D. Medical

Sciences 463, 2009. 93 p.

ISBN 978-951-27-1363-9

ISBN 978-951-27-1380-6 (PDF)

ISSN 1235-0303

ABSTRACT Apoptosis or programmed cell death plays an important role in the regulation of B cell survival.

Failures in the apoptotic signaling can lead to autoimmune diseases or cancer. Moreover, apoptosis

is a major mode of cell death in cancer therapy, and resistance to therapy is often associated with

disturbances in the apoptotic signaling. The detailed knowledge of the apoptotic signaling pathways

can be used as an advantage in the development of new treatment modalities against B cell

malignancies. Two major signaling pathways leading to apoptosis have been described. The

extrinsic pathway is commonly triggered by activation of death receptors, such as Fas/CD95, and

involves the activation of the cysteine protease caspase-8. The intrinsic pathway is initiated by

various extracellular or intracellular stress signals, which induce changes in mitochondrial

membrane permeability and trigger the activation of the caspase-9. The initiator-caspases -8 or -9

activate the cascade of downstream caspases, which are responsible for organized degradation of

cellular organelles.

During adaptive immune response, B cells are selected based on their affinity for foreign antigens,

such as surface molecules of bacteria. The selection of antigen activated B cells takes place at the

germinal centers (GC) of lymphoid organs, and is regulated by surface receptors including B-cell

receptor (BCR), Fas/CD95 and CD40. In this study, BCR- and Fas -induced apoptosis of follicular

lymphoma cells was inhibited by CD40 stimulation. BCR-induced apoptosis involved the

mitochondrial breach and caspase-9 activation pathway, while in Fas-induced apoptosis caspases-8

and -3 were activated independently of mitochondria. The CD40-mediated protection against Fas-

induced apoptosis was associated with a rapid and NF-κB-dependent up-regulation of c-FLIP

proteins, natural inhibitors of caspase-8 activation. Based on our findings, BCR-induced apoptosis

may be involved in the deletion of self-reactive cells generated during the GC reaction, while Fas-

induced apoptosis may be involved in the deletion of low affinity B cells.

Follicular lymphoma is a cancer that originates from GC B cells. The second aim of this study was

to explore signaling pathways of apoptosis induced by rituximab, a chimeric monoclonal antibody

used for the treatment of B cell lymphomas. Rituximab-mediated apoptosis of follicular lymphoma

cells was dependent on the activation of caspase-9 and mitochondrial breach, while the death

receptor pathway was not involved. Moreover, the simultaneous triggering of Fas enhanced

significantly apoptosis induced by rituximab. Thus, the combined use of rituximab with drugs

designed to destabilize mitochondrial membranes, or with drugs that induce the activation of the

death receptor pathway, is warranted for further investigation in clinical settings.

National Library of Medicine Classification: QU 375, WH 200, WH 525, QZ 267

Medical Subject Headings: Apoptosis; Cell Death; Cell Survival; B-Lymphocytes; Signal

Transduction; Caspases; Caspase 8; Caspase 9; Mitochondria; Mitochondrial Membranes;

Receptors, Antigen, B-Cell; Antigens, CD95; Antigens, CD40; NF-kappa B; CASP8 and FADD-

Like Apoptosis Regulating Protein; Antibodies, Monoclonal; Receptors, Death Domain;

Lymphoma, Follicular/drug therapy

ACKNOWLEDGEMENTS

This study was carried out at the Department of Clinical Microbiology, Institute of Clinical

Medicine, University of Kuopio, during the years 2000-2009.

I express my deepest gratitude to my supervisor Professor Jukka Pelkonen, M.D., Ph.D. for his

positivism and encouraging support during this study. I want especially thank Jukka for his wide

expertise on the field of immunology and molecular biology, and always open-minded attitude for

new ideas. I owe my sincere thaks to my second supervisor Mine Eray M.D, Ph.D., for many

supporting discussions, and for her work contribution in the establishment and charachterization of

follicular lymphoma cell lines used in this study.

I address my warmest thanks to all my former and current colleagues at the Department of

Clinical Microbiology for their unforgettable companionship during all these years. Especially, I

want to thank my colleague and dear friend Ulla Nuutinen M.Sc. for her invaluable support, critical

review of the manuscripts and for fascinating discussions. I also wish to thank other members of our

“B cell group” Mikko Mättö Ph.D., Antti Ropponen M.Sc., Ville Postila M.D., and Anna-Riikka

Pietilä M.Sc. for their collaboration and invaluable help during these years. This study would not

exist without your contribution and expertise. My special thanks belong to Pia Keinänen, Päivi

Kivistö, Riitta Korhonen and Eila Pelkonen for their excellent technical assistance and friendship.

I am deeply grateful to official reviewers of this thesis, Professor Mauno Vihinen and Professor

Marko Salmi, for critical evaluation of this thesis and their most valuable comments.

I wish to thank all my friends for their support during the occasional disappointments during this

study, and for sharing many delight running, biking, skiing, paddling, skating and hiking moments

with me.

I owe my sincere thanks to my parents Päivikki and Hannu for giving me support and

encouragement during all these years, and providing me home where education was greatly

appreciated. I also express warm thanks to my sister Salla and her husband Renato, and my sister

Suvi-Tuuli and her life-companion Jaakko for sharing enjoyable moments during our holidays. I am

deeply indebted to Suvi-Tuuli for her precious contribution in the laboratory works, and caring for

Pihla when ever it was needed.

Finally, I dedicate my warmest thanks to Jarno for love, care and sharing the dark and delight

moments during this project. My most loving thanks goes to our little sunshine Pihla, for fulfilling

our days with joy of living, and for allowing me to accomplish this dissertation project during the

maternity leave.

This work was financially supported by Kuopio University Hospital (EVO-fund), Finnish Medical

Foundation, and KELA- the Social Insurance Institute of Finland.

Kuopio, October 2009

Jonna Eeva

ABBREVIATIONS

AIF apoptosis inducing factor

Bax Bcl-2 associated X protein

BAFF B-cell-activating factor of the tumor necrosis factor family

Bcl-2 B-cell lymphoma 2 protein

Bcl-xL Bcl-2 related gene (large variant)

BCR B-cell receptor

BH Bcl-2 homolgy

CD cluster of differentiation

CREB cAMP response element-binding protein,

DISC death-inducing signaling complex

DN dominant negative

ERK extracellular signal regulated kinase

FADD Fas associated death domain

FL follicular lymphoma

FLIP flice inhibitory protein

FLIPI follicular lymphoma prognostic index

GC germinal center

GFP green fluorescence protein

GSK3 glycogen synthase kinase-3

IL interleukine

IAP inhibitor of apoptosis protein

IMS intermembrane space

JNK c-Jun N-terminal kinase

kDa kilo Dalton

lpr lymphoproliferative

mAb monoclonal antibody

MAPK mitogen activated protein kinase

MOMP mitochondrial outer membrane permeabilization

NF-κB nuclear factor-κB

NF-AT nuclear factor of activated T cells

SHM somatic hypermutation

SMAC second mitochondria-derived activator of caspase/direct IAP-binding protein with

low PI

OM outer membrane

PI propidium iodide

PI3K phosphatidyl-inositol-3-kinase

PLC phospholipase-c

PKC protein kinase c

TNF tumor necrosis factor

TRAIL tumor necrosis factor-related apoptosis inducing ligand

TRAF tumor necrosis factor associating factor

zVAD benzyloxycarbonyl-Val-Ala-Asp

Ψm mitochondrial membrane potential

LIST OF ORIGINAL PUBLICATIONS

The thesis is based on the following original publications, which are referred to in the text by the

Roman numerals (I-IV).

I Eeva J *, Postila V*, Mättö M, Nuutinen U, Ropponen A, Eray M, Pelkonen J. Kinetics and

signaling requirements of CD40-mediated protection from B cell receptor-induced apoptosis.

Eur J Immunol 2003;33:2783-91.

II Eeva J, Ropponen A, Nuutinen U, Eeva ST, Mättö M, Eray M, Pelkonen J. The CD40-induced

protection against CD95-mediated apoptosis is associated with a rapid upregulation of anti-

apoptotic c-FLIP.

Mol Immunol 2007;44:1230-7

III Eeva J, Nuutinen U, Ropponen A, Mättö M, Eray M, Pellinen R, Wahlfors J, Pelkonen J.

Feedback regulation of mitochondria by caspase-9 in the B cell receptor-mediated apoptosis.

Accepted for publication. Scand J Immunol.(In press)

IV Eeva J, Nuutinen U, Ropponen A, Mättö M, Eray M, Pellinen R, Wahlfors J, Pelkonen

J. The involvement of mitochondria and the caspase-9 activation pathway in rituximab-

induced apoptosis in FL cells. Apoptosis 2009;14:687-98.

* Equal contributors

The original publications have been reproduced with permission of the copyright holders.

The thesis includes also previously unpublished data.

CONTENTS

1. INTRODUCTION ............................................................................................. 13

2. REVIEW OF THE LITERATURE ................................................................ 15

2.1. Apoptosis ..................................................................................................... 15 2.1.1 Different ways to die; apoptosis, autophagy and necrosis ............................................... 15 2.1.2 Mitochondrial changes during apoptosis ......................................................................... 17 2.1.3 Caspases – executioners of the apoptotic cell death ........................................................ 22

2.2. B cells and the humoral immune response .............................................. 24 2.2.1 An overview of the humoral immune response ............................................................... 24 2.2.2 The B cell antigen receptor; structure and activation ...................................................... 25 2.2.3. B cell maturation ............................................................................................................ 28 2.2.4 Life and death decisions of a B cell; regulation of B cell apoptosis ................................ 32 2.2.5 GC as an origin of B cell malignancies ........................................................................... 35

2.3. Monoclonal antibody therapy for cancer treatment ............................... 37

2.4. Follicular lymphoma .................................................................................. 40 2.4.1. Incidence and clinical characteristics ............................................................................. 40 2.4.2 Pathophysiology .............................................................................................................. 40 2.4.3 Current treatment of follicular lymphoma ....................................................................... 41

3. AIMS OF THE STUDY .................................................................................... 46

4. MATERIALS AND METHODS ..................................................................... 47

4.1. Cell line and culturing ............................................................................... 47

4.2. Establishment of overexpressing cell lines ............................................... 47

4.3. Cell treatment experiments ....................................................................... 48

4.4 Apoptosis detection ..................................................................................... 49 4.4.1 Quantification of fractional DNA content (I-IV) ............................................................ 49 4.4.2 Mitochondrial membrane potential depolarization (I-IV) ............................................... 50 4.4.3 Cytochrome c release (I-IV) ............................................................................................ 50 4.4.4 Caspase activation (I-IV) ................................................................................................ 50 4.4.5 Cell membrane permeabilization III ................................................................................ 51

4.5. Western blotting ......................................................................................... 51

4.6. Reverse transcriptional (RT)-polymerase chain reaction ...................... 52

4.7 Statistical analyses....................................................................................... 52

5. RESULTS .......................................................................................................... 53

5.1. The molecular mechanisms of B cell receptor-induced apoptosis (I, III)

............................................................................................................................. 53 5.1.1 The signal transduction pathways connected with B cell receptor-induced apoptosis .... 53 5.1.2 The role of caspase-9 and mitochondrion in B cell receptor-induced apoptosis (I, III) .. 54 5.1.3 The caspase-8 activation pathway showed only marginal role in B cell receptor-induced

apoptosis................................................................................................................................... 56 5.2. The molecular mechanisms of Fas/CD95-induced apoptosis (II, III, IV)

............................................................................................................................. 58

5.3. The molecular mechanisms of rituximab-induced apoptosis (IV) ......... 58

5.4. Molecular mechanisms of CD40-mediated protection against apoptosis

(I, II and IV) ...................................................................................................... 60 5.4.1 Kinetics and signaling requirements of CD40 mediated protection from B cell receptor-

mediated apoptosis (I) .............................................................................................................. 61 5.4.2 The CD40-induced protection against Fas-induced apoptosis was associated with a rapid

up-regulation of anti-apoptotic c-FLIP (II) .............................................................................. 61 6. DISCUSSION .................................................................................................... 63

6.1. FL cell lines as an experimental model for the study of receptor-

mediated apoptosis ............................................................................................ 63

6.2. Molecular mechanisms of receptor-mediated apoptosis ........................ 64 6.2.1 Fas-induced apoptosis proceed via caspase-8 activation pathway while B cell receptor-

and rituximab –induced apoptosis are dependent on mitochondrial pathway .......................... 64 6.2.2 Feedback regulation of mitochondria by caspase-9 during intrinsic apoptosis ............... 65 6.2.3. Caspase-9 –mediated apoptosis was poorly inhibited by pan-caspase-inhibitor z-VAD-

fmk ........................................................................................................................................... 67 6.2.4. What happens before mitochondrial breach during intrinsic apoptosis? ........................ 67

6.3. Molecular mechansims of CD40-mediated protection against apoptosis

(I, II and IV) ...................................................................................................... 68

6.4. A proposed model of the molecular interactions during the germinal

center negative selection ................................................................................... 71

7. CONCLUSIONS; Therapeutic applications and future directions ............. 73

7.1. Cell surface receptors in the regulation of life and death decisions in the

germinal center .................................................................................................. 73

7.2. CD40-CD40L interactions as potential targets for cancer therapy ....... 74

7.3. Novel approaches to enhance rituximab therapy.................................... 75

8. REFERENCES .................................................................................................. 77

APPENDIX: ORIGINAL PUBLICATIONS I-IV

13

1. INTRODUCTION

In our body, millions of cells are being produced every day. To maintain homeostasis, the

divisions of cells must be delicately balanced by deletion of unnecessary cells. Elegantly, these cells

are cleared by a tightly regulated process called apoptosis or programmed cell death.

Apoptosis is genetically regulated form of cell death, in which a single cell is sacrificed for the

benefit for the whole organisms. There are several circumstances in which multicellular organisms

benefit for the targeted deletion of cells. For example, cells irreversibly injured by virus infection,

radiation, growth factor withdrawal or other violating signals from outside of the cell are no longer

useful for the organism and thus eliminated. In addition, various stress signals originating from

inside of the cell, such as DNA lesions that have occurred during cell divisions, can induce

programmed cell death. Unfortunately, some of the cells which have gained abnormalities in genes

regulating apoptosis may escape from cell death and continue to proliferate. These cells are

potentially hazardous, since the further accumulation of growth promoting and apoptosis inhibiting

lesions can in rare circumstances lead to development of cancer, a disease characterized by

uncontrolled cell proliferation.

Follicular lymphoma (FL) is a heterogeneous cancer disease of B cell origin. In Finland, over two

hundred follicular lymphoma cases are diagnosed annually. Until these days, this disease has been

assumed as incurable, and thus only the patients with symptoms or with advanced disease are

treated with standard chemotherapy. However, immunotherapy with anti-CD20 antibody rituximab

has recently improved the overall survival and disease free time of FL patients when combined to

standard chemotherapy. When given to patients, rituximab depletes B cells by three overlapping

mechanisms; complement-mediated cytotoxicity, antibody-dependent cytotoxicity and apoptosis.

Thus far, the molecular mechanisms of rituximab-induced apoptosis are largely unsolved. The great

advantage of rituximab and also other therapies based on the use of immunotherapeutic antibodies is

their specificity for the cancer cells and thus milder side effects as compared to standard

chemotherapies. Inspired by rituximab, several immunotherapeutic drugs are currently under

clinical or preclinical trials for the treatment of B cell lymphomas and also for other cancer diseases.

Study of apoptotic signaling pathways is important, since these pathways are potential targets

when developing new treatment modalities for cancer. Commonly, cancer is associated with

dysregulation of apoptotic pathways, which offers targets for drug discovery. Majority of cancer

therapeutics used currently delete cancer cells by apoptosis. Thus, failures in apoptotic signaling can

14

result in the resistance against cancer therapy, which may be hindered by drugs promoting

apoptosis. This work was aimed to reveal the molecular mechanisms involved in the receptor-

mediated apoptosis of B cells. Apoptotic pathways induced by B cell receptor and Fas are important

regulators of B cell fate during B cell development and immune reactions. The molecular

mechanisms of CD20-induced cell death are currently unknown despite the wide use of anti-CD20

antibody rituximab in the immunotherapy. This work proposes signaling pathways involved in

BCR, Fas and CD20 –mediated apoptosis in a FL cell line. The detailed knowledge of the apoptotic

signaling pathways may be used in the future for the development of novel, apoptosis directed

treatment modalities against cancer diseases.

15

2. REVIEW OF THE LITERATURE

2.1. Apoptosis

2.1.1 Different ways to die; apoptosis, autophagy and necrosis

In the human body dazzling rate of continuous cell division must be compensated by cell death to

maintain tissue homeostasis. Either deficient or excessive cell death can lead to a variety of

pathologic conditions. For example, insufficient death of cells with deleterious genetic lesions may

result in the development of cancer. Conversely, ischemia induced cell death during stroke or

myocardial infarction incurs irreversible tissue damage.

Cell death can be classified according to morphological appearance, biochemical patterns and

functional aspects (programmed versus accidental) (Galluzzi et al., 2007). Mainly based on the

morphological features, three different forms of cell death can be classified; apoptosis, autophagy

and necrosis. However, in some instances the dying cell has overlapping features of different cell

death modalities making the exact classification unfeasible.

Apoptosis or programmed cell death (type I cell death) is genetically engineered, physiological

form of cell death where the single cell is sacrificed for the benefit for the whole organism. The

term apoptosis was first introduced in the 1970`s century by Kerr and Wyllie, who first described

the cell death with specific morphological features (Kerr et al., 1972). These well defined

morphological changes include cell shrinkage and nuclear condensation (pyknosis), nuclear

fragmentation (karyohexis) and plasma membrane blebbing (Galluzzi et al., 2007). Finally the intact

cytoplasmic organelles or fragments of the nucleus are cleanly packaged inside membrane bounded

vacuoles which are also called apoptotic bodies. Eventually, these apoptotic bodies are recognized

and engulfed by phagocytic cells, without evoking an inflammatory response which could be

harmful for the surrounding tissue. Apoptotic cell death is most often associated with typical

biochemical events including DNA fragmentation, cysteine protease activation and mitochondrial

membrane potential collapse, but these events are not definitive for apoptosis (Galluzzi et al., 2007).

Autophagy (type II cell death) is a process where parts of the cytoplasm are sequestered within

double-membrane vacuoles and finally digested by lysosomal hydrolases (Kroemer and Jäättelä,

2005). Autodigestion of cellular constituents may serve as a defense mechanism against nutrient

deficiency or sub-lethal cellular injury and therefore autophagy may even prevent the cell death in

some conditions. However, extensive autodigestion may lead to cell death especially in the cases

16

where apoptosis is somehow disturbed. The morphological hallmarks of autophagy include the

double-membrane autophagic vacuoles and the lack of chromatin condensation and other typical

features of apoptosis (Galluzzi et al., 2007).

Necrosis (type III cell death) is usually defined as a harmful event as it is often associated with

pathological conditions involving excessive cell loss, and the disadvantage of necrotic cells to

promote inflammatory response (Galluzzi et al., 2007). The necrotic cells undergo extensive

swelling and unorganized dismantling of cytoplasmic organelles which eventually leads to the

rupture of plasma membrane and spilling of the cell contents to surrounding extracellular space.

Necrosis lacks the typical morphological characteristics of apoptosis and massive vacuolization seen

during autophagic cell death. Necrosis is most often accompanied with the mitochondrial

dysfunction including extensive production of reactive oxygen species, mitochondrial membrane

permeabilization (MMP), ATP depletion and failure in Ca2+

homeostasis. The activation of calpain

and cathepsin proteases and the lysosomal rupture are typically observed during the necrotic cell

death. While sufficient ATP and glucose content are obligatory for apoptosis, their depletion favors

switching to necrosis. However, the view that apoptosis is the only physiological or advantageous

form of cell death whereas necrosis is always harmful and pathological process has recently gained

a momentum. Indeed, apoptotic cell death can be shifted to necrosis in the circumstances where the

activation of caspase-activation is artificially blocked (Goldstein et al., 2005). In addition, in some

instances necrosis can be programmed and triggered appropriately by specific plasma membrane

receptors, although these characteristics of cell death have been previously associated only with

apoptotic cell death (Galluzzi et al., 2007). Thus, there is marked overlap between the different cell

deaths, and in many cases features of different forms of cell death can be observed simultaneously.

The next chapters are focused on the detailed examination of apoptotic signaling pathways. In

general, there are two main apoptotic pathways, the extrinsic and intrinsic pathways, which differ in

their inducing agent and the involvement of mitochondria (Figure 1.). Both of these pathways lead

to the activation of a series of cysteine proteases called caspases, which are activated in a cascade in

which activated caspases cleave and activate downstream caspases (Boatright and Salvesen, 2003).

The extrinsic pathway is commonly triggered by activation of death receptors, such as Fas/CD95,

and involves the activation of the cysteine protease caspase-8. The intrinsic pathway is initiated by

various extracellular or intracellular stress signals, which induce changes in mitochondrial

membrane permeability and trigger the activation of the caspase-9. In both pathways, the initiator-

17

caspases activate the cascade of caspases, which are responsible for proteolysis of the cell

constituents.

Figure 1. The extrinsic and intrinsic apoptotic signaling pathways

The extrinsic pathway is initiated by ligation of death receptors belonging to TNF-receptor

superfamily (TNFR1, Fas/CD95, TRAIL-R1/R2). The intrinsic pathway is triggered as a result of

stress signals that converge on the level of mitochondria. Both pathways result on the activation of

executor caspases (caspase-3 and -7) which are responsible for the targeted proteolysis during the

apoptotic cell death.

2.1.2 Mitochondrial changes during apoptosis

Mitochondria are essential for life as they are responsible for oxidative energy production. In

addition, mitochondria are centrally involved in the regulation of cell death. Mitochondria contain a

highly folded inner membrane (IM) which is involved in the ion transport and energy production

and an outer membrane (OM) which sequesters the intermembrane space (IMS) from the cytosol.

Stress stimuli

Bax/Bak

Bcl-2 Bcl-XL

Mitochondria

Apoptosis

FLIP Apaf Cyt-c Cyt-c

Smac

HtrA2

IAPs

Death receptor

A) Extrinsic pathway B) Intrinsic pathway

FADD

Caspase-8

Bid

Caspase-3

Caspase-9

18

Variety of lethal signals from the endoplasmic reticulum, cell surface receptors, nucleus or

extracellular environment converges on mitochondria to induce the mitochondrial outer membrane

permeabilization (MOMP). As a result of MOMP, caspase-activating molecules, such as

cytochrome c, and caspase-independent death effectors are released from the intermembrane space

of mitochondria to the cytosol (Figure 1.). The permeabilization of the outer mitochondrial

membrane is frequently associated with the permeabilization of the inner mitochondrial membrane,

which leads to the collapse of mitochondrial membrane potential (ΔΨm), metabolic failure and

eventually cell death.

The permeabilization of mitochondrial outer membrane is regulated by Bcl-2 family proteins and

is considered as “point of no return” in apoptosis (Garrido et al., 2006; Green and Kroemer, 2004).

The Bcl-2 family proteins are divided into two main groups according to their capacity to either

inhibit or promote apoptosis (Kuwana and Newmeyer, 2003) (Figure 2.). The pro-apoptotic family

members are further classified according to whether they have multiple Bcl-2 homology (BH)

domains or only one BH domain (BH3-only proteins). The common feature of Bcl-2 family proteins

is the formation of homo- or heterodimers with other family members thus neutralizing the action of

each other.

Several pieces of evidence demonstrate that MOMP is a central event during apoptosis, and that

the permeabilization process is precisely regulated by the Bcl-2 family proteins. Firstly, anti-

apoptotic Bcl-2 and Bcl-xL block the release of pro-apoptotic molecules from the intermembrane

space and thus inhibit apoptosis (Kluck et al., 1997; Kuwana et al., 2002; Yang et al., 1997).

Secondly, the pro-apoptotic Bcl-2 family members can permeabilize lipid bilayers, allowing the

release of macromolecules across the membrane. The first clues that the Bcl-2 family proteins can

permeabilize lipid membranes come from their structure, which has similarities with pore forming

bacterial toxins (Suzuki et al., 2000). Afterwards, it has been shown that oligomerized Bax can

form channels in plain liposomes through which cytochrome c could be released (Saito et al., 2000).

Moreover, it has been demonstrated that BH3 only protein Bid activates Bax to produce membrane

openings in mitochondrial outer membrane large enough to transmit mitochondrial IMS proteins

(Kuwana et al., 2002). This process could be inhibited by Bcl-xL and involved cardiolipin, an

abundant lipid in the mitochondrial membrane. These findings were reproduced in cell models since

it was found that cells lacking both Bax and Bak were completely resistant to cytochrome c release

and apoptosis induced by activated Bid (Wei et al., 2001). Thus, it is now widely accepted that

19

during apoptosis, Bax and Bak oligomerize and insert on the mitochondrial outer membrane

resulting in its permeabilization and the release of pro-apoptotic proteins from IMS to cytosol.

Mitochondrial outer membrane permeabilization (MOMP) is regulated by the Bcl-2 family

proteins

Figure 2. The Bcl-2 family of proteins

The Bcl-2 family of proteins is divided on three groups based on their composition of Bcl-2

homology (BH) domains. The anti-apoptotic members contain four domains (BH1-4). The pro-

apoptotic members are divided in to two groups; multidomain proteins (Bax, Bak) which contain

three domains (BH1-3), and the BH3-only proteins. Often, the BH3-only are subdivided into direct

activators (Bid, BIM) and de-repressors/ sensitizers. Each protein contains the hydrophobic

carboxyl terminal transmembrane domain (TM).

Inactive Bax exist in the cytosol as monomers whereas Bak is constitutively bound in

mitochondrial outer membranes (Hsu and Youle, 1998; Youle and Strasser, 2008). It is generally

believed that the activation of Bax/Bak involves the translocation of Bax to outer membrane and

oligomerization with Bak to form membrane spanning pores. The activation of Bax/Bak is induced

either directly or indirectly by “activator” BH3-only proteins, including Bid and Bim. According to

“direct activation model”, BH3-only proteins (i.e. Bid and Bim) bind directly to Bax/Bak leading to

their activation (Kim et al., 2006). According to an alternative “indirect activation model”, BH3-

Anti-apoptotic Bcl-2 proteins

Pro-apoptotic Bcl-2 proteins Multiple BH3 domains

BH3-only

Bcl-2, Bcl-w, Bcl-XL, Mcl-1, A1

Bax, Bak

Bid, Bim

Bad, Bik, Bmf, Hrk, Noxa, Puma

BH4 BH3 BH1 BH2

BH3 BH1 BH2

TM

TM

TM BH3

20

only proteins promote apoptosis exclusively by binding and neutralizing anti-apoptotic Bcl-2 family

members and thus unleashing Bax and Bak from their control. This model is supported by the

findings that Bax and Bak do not bind to “activating” BH3-only proteins, and apoptosis proceed

normally in cells deficient of Bim and Bid (Willis et al., 2007). The interactions between different

Bcl-2 family members are still not fully elucidated and different set of these proteins depending on

the cell type or an apoptosis inducing agent may be involved.

The mitochondrial permeability transition and the membrane potential collapse during

apoptosis

The mitochondrial permeability transition (MPT) denotes the increase of the permeability of the

mitochondrial inner membrane that leads to the collapse of ΔΨm, mitochondrial swelling, and

rupture of the outer mitochondrial membrane (Tsujimoto and Shimizu, 2007). It is generally thought

that MPT is induced as a result of channel formation (permeability transition pore, PTP) between

the inner and outer membrane. This channel is thought to consist of the voltage-dependent anion

channel (VDAC) in the outer membrane, the adenine nucleotide traslocator (ANT) in the inner

membrane and the cyclophilin D (Cyp D) in the matrix.

The role of MPT in apoptosis is controversially discussed. The hypothesis that MPT is involved in

apoptosis is largely based on findings that inhibitors of Cyp D or ANT, cyclosporine A and

Bonkrecik acid respectively, can inhibit apoptosis in some experimental models (Custodio et al.,

2001; Zamzami et al., 1996). In addition, MPT may result in the permeability changes of the outer

mitochondrial membrane leading to the cytochrome c release during apoptosis (Green and Kroemer,

2004). However, recently it has been demonstrated that cell lines isolated from Cyp D deficient

mouse undergo apoptosis normally in response to various stimuli, showing that MPT (or at least its

component Cyp D) is not involved in apoptosis (Nakagawa et al., 2005). In contrast, Cyp D and

MPT seem to be instrumental in some forms of necrotic cell death (Nakagawa et al., 2005).

However, the collapse of ΔΨm is frequently seen in apoptosis induced by variety of stimuli. Thus it

can be speculated that the permeabilization of the inner membrane participates also to apoptotic cell

death. The ΔΨm collapse leads to cessation of ATP synthesis, release of Ca2+ from the matrix,

production of reactive oxygen species (Skommer et al., 2007), and these changes may promote both

apoptotic and nectrotic cell death.

21

Cytochrome c and other pro-apoptotic molecules released from the IMS

Cytochrome c is a small heme-containing protein bounded to outer face of the IM and involved in

the electron transport and oxidative phosphorylation (Garrido et al., 2006). In addition to its vital

function in the respiratory chain, cytochrome c has distinct function as a regulator of apoptosis.

Upon apoptotic stimulus, cytochrome c is released through permeabilized mitochondrial outer

membrane in to the cytosol, where it binds to apoptotic protease-activating factor-1 (Apaf-1) and

induces ATP dependent formation of protein complex called the apoptosome (Riedl and Salvesen,

2007). In addition to cyt c, ATP and Apaf-1, cysteine protease caspase-9 is recruited to apoptosome

where it undergoes dimerization and activation. Caspase-9 functions as an initiator caspase of the

mitochondrial pathway of apoptosis and further activates the executioner caspases responsible for

the targeted proteolysis of cellular substrates (see later).

Recently, Hao et al. developed a mouse model in which mutated cytochrome c retained normal

respiratory function but lacked apoptotic function due to failure to oligomerize with Apaf-1 (Hao et

al., 2005). In these animals, apoptosis during brain- and lymphocyte development was severely

disrupted showing the importance of cytochrome-c mediated apoptosis in the organogenesis.

Although the fibroblasts of this cytocrome-c mutated mouse model were resistant to caspase

activation and apoptosis, the thymocytes were sensitive to apoptosis induced by variety of signals.

Thus the authors concluded that there may be cytochrome-c / apoptosome independent pathways of

caspase-activation. Indeed, in addition to cytochrome c, also other proapoptotic molecules are

released from the IMS during apoptosis. The SMAC/DIABLO and HtrA2/Omi released to the

cytosol bind and inactivate IAPs (Inhibitor of Apoptosis Proteins) and thus facilitate the activation

of caspases (Figure 1.) (Suzuki et al., 2001; Verhagen et al., 2000). Apoptosis-inducing factor (AIF)

is expelled from IMS due to apoptotic stimulus and translocates to nucleus where it contributes to

DNA fragmentation (Porter and Urbano, 2006). However, AIF is centrally involved in the oxidative

phosphorylation, and thus it has been speculated that the main mechanism of how AIF release

triggers cell death is related to failure in energy metabolism (Porter and Urbano, 2006). Endo G was

previously shown to cause chromatin condensation and DNA fragmentation independently of

caspase activation. However, the role of EndoG in apoptosis is recently questioned in the basis of

genetic studies (Irvine et al., 2005).

22

2.1.3 Caspases – executioners of the apoptotic cell death

Caspases are cysteine proteases which play an important role in the regulation and execution of

apoptotic cell death. The term caspase is derived from cysteine dependent, aspartatic acid specific

protease, which denotes that the cysteine is involved in the catalytic site of the protease which

cleaves its substrate proteins after the aspartatic acid residue (Boatright and Salvesen, 2003).

Recognition of at least four amino acids next to the cleavage site is involved for substrate binding

and cleavage (Thornberry and Lazebnik, 1998). This four amino acid recognition sequence differs

among the caspases and explains the differences in their action. The activation of caspases does not

lead to indiscriminate protein digestion, instead highly specific set of target proteins is cleaved

elegantly resulting in the loss or change in their function.

The caspases are synthesized and stored as inactive precursors (zymogens) which became

activated upon apoptosis induction (Boatright and Salvesen, 2003). The first caspases to be

activated in the proteolytic cascade are termed initiator caspases (caspases -2, -8, -9 and -10). The

zymogens of the initiator caspases exists within a cell as inactive monomers, which became

activated by dimerization at multiprotein activating complexes (Boatright and Salvesen, 2003;

Boatright et al., 2003; Donepudi et al., 2003). In general, two ways of initiator caspase-activation

have been described; the extrinsic pathway is initiated by ligation of death receptors belonging to

TNF-receptor superfamily (TNFR1, Fas/CD95, TRAIL-R1/R2) and the intrinsic pathway is

triggered by MOMP as a result of stress signals that converge on the level of mitochondria (Figure

1.).

The death receptor (extrinsic) pathway

The extrinsic pathway of apoptosis is triggered by binding of extracellular ligand (TNF, FAS-L,

TRAIL) to their specific transmembrane receptors (Ashkenazi and Dixit, 1999). However, in the

case of Fas/CD95 receptor, also ligand independent activation by receptor aggregation has been

described (Hennino et al., 2001). Upon activation, the death receptors form trimeric complexes

which bind and recruit the adaptor protein FADD (Fas-associated death domain). FADD in turn

binds caspase-8 zymogens resulting in the formation of the death inducing signaling complex

(DISC) (Medema et al., 1997), a multiprotein complex which promotes the dimerization of caspase-

8 leading to its activation (Donepudi et al., 2003). In addition to caspase-8, caspase-10 can be

recruited to DISC and work as an initiator caspase of the death receptor pathway at least in some

cell models (Kischkel et al., 2001).

23

Two types of cells which differ in their involvement of mitochondria in the death receptor

mediated apoptosis has been described (Scaffidi et al., 1998). In type I cells the amount of caspase-8

activated at the DISC is sufficient for the activation of the executioner caspases without the

involvement of mitochondria. In type II cells, small amount of activated caspase-8 cleaves and

activates Bid, a pro-apoptotic member of the Bcl-2 family, leading to its translocation to

mitochondria where it induces MOMP and engages the extrinsic pathway with the intrinsic pathway

of apoptosis (Li et al., 1998) (Figure 1.).

Besides functioning as an activation platform for caspase-activation, DISC is also involved in the

regulation of the initiator caspase activation by recruiting the caspase-8 homolog FLIP (FLICE-like

inhibitory protein) (Irmler et al., 1997; Scaffidi et al., 1999). FLIP has structural homology with the

caspase-8 and can thus replace it in the DISC. Generally, FLIP is considered as an apoptosis

inhibiting molecule as it can displace caspase-8 from the DISC while it does not itself have notable

proteolytic activity.

The mitochondria-dependent (intrinsic) pathway

The intrinsic or mitochondrial pathway of apoptosis is triggered by multiple cellular or

extracellular stress signals, such as growth factor withdrawal, DNA damage or radiation, which

converge on the level of mitochondria. The hallmark of the intrinsic pathway is the activation of

caspase-9 at the multiprotein activating complex called the apoptosome (Riedl and Salvesen, 2007).

As a result of mitochondrial permeability change, cytochrome c is released into the cytosol, where it

binds to an intracellular receptor molecule apoptosis-inducing factor-1 (Apaf-1) leading to its

oligomerization and ATP dependent conformational change. Apaf-1 in turn recruits caspase-9 via its

N-terminal caspase-activation recruitment domain (CARD). Activation of caspase-9 monomers at

the apoptosome is achieved by dimerization (Riedl and Salvesen, 2007). Moreover, it has been

suggested that the apoptosome increases the caspase-9 affinity for procaspase-3 (Yin et al., 2006)

Activation of the executioner caspases (caspases -3, -6 and -7) involves their proteolytic

processing by the initiator caspases (Thornberry and Lazebnik, 1998). The cleavage of the linker

domain of executioner caspases results in the formation of small and large caspase subunits, which

heterodimerize to form a catalytical unit. Executioner caspases contribute to apoptosis by cleaving

structural proteins of the nucleus (lamin), proteins involved in the regulation of cytoskeleton

(gelsolin, focal adhesion kinase) and proteins involved in the DNA repair (DNA-PKcs) or replication

(replication factor C) (Thornberry and Lazebnik, 1998). In addition, executioner caspases can

24

enhance mitochondrial dysfunction and hence cell death by cleavage of the respiratory chain

components (Ricci et al., 2004) or anti-apoptotic Bcl-2 family proteins (Chen et al., 2007; Cheng et

al., 1997).

The activity of executioner caspases is kept in check by Inhibitor of Apoptosis Proteins (IAPs) to

ensure that these proteases are not activated inappropriately (Verhagen et al., 2001). To date, several

members of the IAP-family have been characterized, including XIAP, cIAP1, cIAP2 and surviving

(reviewed in (D'Amelio et al., 2008)) The IAPs in turn can be inactivated by SMAC/DIABLO or

HtrA2/Omi released from IMS during the apoptotic insult (Figure 1.) (Suzuki et al., 2001; Verhagen

et al., 2000; Verhagen et al., 2001).

In addition of their death promoting effect during apoptosis, caspases may also function in the

regulation of cell survival, proliferation, differentiation and inflammation (Lamkanfi et al., 2007).

2.2. B cells and the humoral immune response

2.2.1 An overview of the humoral immune response

The main function of the immune system is to protect our body from invading infectious agents.

The adaptive immune response enables the specific destruction of invading micro-organisms such

as viruses and bacteria, whose foreign structures work as antigens. The immunological memory and

the adaptation to combat micro-organisms that are transforming all the time are instrumental

properties of the adaptive immunity.

Failures in the immune system are associated with several diseases. Disturbance in the

development of a specific immune response (immunodeficiency) leads to overwhelming or

recurrent infections. On the other hand, a normal immune response can be inappropriately directed

against harmless, non-infectious agents, such as in the case of allergy or autoimmune diseases.

T and B lymphocytes are greatly in charge for the development of a specific immune response. Each

individual lymphocyte bears a unique variant of an antigen receptor, so that lymphocytes

collectively have an enormous repertoire of antigen receptors which can bind highly diverse

antigens.

The humoral immune response is mediated by antibodies (immunoglobulins), which are secreted

by terminally differentiated B cells called plasma cells. The antibody is a soluble form of the B cell

surface antigen receptor, and recognizes the same antigen as the receptor. Antibodies secreted to

circulation delete micro-organisms by three main ways. Antibodies can bind to micro-organisms

25

and thus prevent their adherence to host cell (neutralization). Secondly, antibodies can bind to the

surface of pathogens and promote its phagosytosis by macrophages and other effector cells

(opsonization). Thirdly, antibody binding to the surface of micro-organism may trigger the

activation of complement system leading to pore formation on the target cell surface and cell lysis.

2.2.2 The B cell antigen receptor; structure and activation

Figure 3. The structure of B cell receptor complex

The surface immunoglobulin consists of two heavy and two light chains combined to each other by

disulfide bonds. Surface immunoglobulin consists of variable and constant regions which determine

the antigen binding capacity and effector functions of the immunoglobulin, respectively. Moreover,

immunoglobulin can be divided on two structural fragments; fragment antigen binding (Fab) and

fragment crystallizable (Fc). The accessory proteins Igα and Igβ contain the ITAM motifs critically

involved in the BCR signaling.

Variable region

Constant region

- Light chain

- Heavy chain

Antigen

Fab

Fc

ITAM P P

Igα Igβ

26

The main functions of the BCR are to recognize foreign antigens, to internalize antigens for

further processing and presentation to T-helper cells, and to transmit activating signals to the cells

interior. The activating signals initiated after BCR-stimulation have different outcomes depending

on the maturational stage of a B cell. In mature B cells, BCR triggering leads to proliferation and

survival, whereas in immature B cells BCR activation may lead to inactivation or apoptosis (Niiro

and Clark, 2002). The central question in the research concerning the BCR signaling is how the

signals initiated from the same receptor can evoke these different responses.

The B cell receptor complex is made up of the cell surface immunoglobulin combined with

accessory proteins Igα and Igβ (Figure 3.) (Reth and Wienands, 1997). The immunoglobulin part

consists of two heavy and two light chains combined to each other by disulfide bonds. The variable

regions of both light and heavy chains are unique in each B cell clone and responsible for antigen

binding specificity. Diversity of the antigen binding capacity is a result of immunoglobulin variable

region gene rearrangements during early B cell development (Busslinger, 2004). The surface

immunoglobulin has the same antigen specificity as the soluble immunoglobulin that the cell will

eventually produce as a plasma cell.

The surface immunoglobulin has a short cytoplasmic tail and can not transmit signals in to cells

interior. Accessory proteins Igα and Igβ provide the cytoplasmic domains crucial for BCR signaling

(Reth and Wienands, 1997). They contain amino acid sequences called immunoreceptor tyrosine-

based activation motifs (ITAMs), which become phosphorylated by Src-family protein tyrosine

kinases (PTKs), such as Lyn, after engagement of BCR by antigen (Niiro and Clark, 2002). The

phosphorylated ITAMs in turn recruit and facilitate the activation of protein kinases Syk and Btk,

which in turn activate the phopholipase Cγ2 (PLC γ2), phosphatidyl-inositol-3-kinase (PI3K) and

Ras-Raf-1-ERK downstream signaling pathways (Figure 4.) (Niiro and Clark, 2002). Adaptor

proteins such as B-cell linker (BLNK) and BAM 32 (B-lymphocyte adaptor molecule of 32 kD)

connect the initial kinases with downstream signaling molecules.

The activation of ERK1/2 (extracellular signal regulated kinase) has been associated with both

proliferation and apoptosis of B cells (Koncz et al., 2002; Lee and Koretzky, 1998). ERK1/2 are

activated by a cascade of upstream protein kinases Ras and Raf-1 (Niiro and Clark, 2002). In

addition to Ras-Raf-1, the PLC-γ2 signaling pathway has been shown to participate in the activation

of ERKs (Hashimoto et al., 1998). In mature B cells, sustained activation of ERK after BCR

stimulation was associated with the up-regulation CREB and Elk-1, transcription factors involved in

cell proliferation (Koncz et al., 2002). In line with this, the pharmacological inhibition of ERK

27

activation suppressed BCR-induced proliferation in mature B cells (Richards et al., 2001). In

contrast, it was shown in the murine immature B cell line that ERK2 plays an active role in anti-

IgM-mediated apoptosis (Lee and Koretzky, 1998). In human B lymphoma cell lines, the Ras-ERK

pathway activation was connected with up-regulation of Bim and induction of apoptosis (Stang et

al., 2009). Thus, the role of the Ras-Raf-1-ERK pathway in the regulation of B cell fate may depend

on the maturational stage of the B cell and the kinetics of ERK activation.

Figure 4. B cell receptor-induced signal-transduction pathways

Three main pathways are activated after B cell receptor activation; Ras, PLC and PI3-kinase

pathways. The most important downstream mediators include mitogen activated protein kinases

(MAPKs), NF-AT, cyclin D and NF-κB.

The PI3K mediates B cell survival by activating downstream kinase AKT which promotes cell

survival by directly inhibiting pro-apoptotic Bcl-2 family member BAD and by inducing the

RAS

Raf

PLC

PKC Ca2+

PI3K

AKT

GSK3

Cyclin D NF-κB NFAT CREB

ERK JNK P38

BCR

MAPKs

Src-family PTKs

28

expression of pro-survival proteins (Bcl- xL, Bcl-2, A1) through activation of NF-κB. In addition,

AKT inhibits GSK3 (glycogen synthase kinase-3) leading to stabilization of Cyclin D and Myc,

thereby facilitating cell proliferation (Niiro and Clark, 2002).

Activated PLCγ cleaves phosphatidylinositol 4,5-bisphosphate (PIP2) to inositol 1,4,5-trisphosphate

(IP3) and diacylglyserol (DAG) (Campbell, 1999). IP3 binds to its receptors on endoplasmic

reticulum leading to the release of intracellular calcium stores, while DAG activates certain

isoforms of PKC. The release of intracellular Ca2+

and PKC activation are crucial for the activation

of mitogen activated protein kinases (MAPKs), such as ERK, JNK and P38 (Niiro and Clark, 2002).

In addition, elevated cytosolic Ca2+

level triggers the activation of calcineurin, a protein

phosphatase, which can activate target molecules including caspase-2, transcription factor NFATc2

or mitogen activated protein kinases (MAPKs) p38 and JNK (Chen et al., 1999; Graves et al., 1996;

Kondo et al., 2003). The activation of JNK and p38 have been associated with BCR apoptosis in

B104 cell line (Graves et al., 1996). In addition, JNK and p38 are activated after various stress

signals, such as radiation induced DNA damage (Chen et al., 1996).

2.2.3. B cell maturation

The selection of immature and transitional B cells; the negative selection during B cell

development in the bone marrow and periphery

During early B cell development in the bone marrow, the B cell precursors undergo recombination

of the immunoglobulin heavy (H) and light (L) chain genes which creates diversity to antigen

binding capacity. The recombination is accomplished by RAG1/2 proteins which rearrange first

variable (V), diversity (D) and joining (J) regions of the Ig heavy chain followed by VJ regions of

the light chain (Busslinger, 2004). If the rearrangement is completed successfully, the B cell

expresses functional surface IgM receptor combined with accessory proteins Igα and Igβ and its

development can continue.

The gene rearrangement process may probably produce B cells with self-reactive receptors which

have to be eliminated to maintain immune system self-tolerant. The mechanisms of this called

negative selection include deletion, receptor editing, anergy and inergy (Hartley et al., 1991).

Deletion of self-reactive immature B cells has been demonstrated in transgenic animal models

(Hartley et al., 1991; Nemazee and Burki, 1989). Moreover, in vitro, immature B cells die by

apoptosis after antigen receptor activation, and these cells have been used as a model of negative

selection (Norvell et al., 1995). Especially, strong signal generated by multivalent antigen drives

29

cells to deletion. However, the bone marrow developing B cells might be rescued by receptor

editing process, whereby self-reactive antigen receptor is replaced by newly generated non-self-

reactive antigen receptor after recurrent immunoglobulin gene-rearrangement (Melamed and

Nemazee, 1997).

Immature B cells which bind soluble self antigens are not deleted immediately in the bone

marrow, but instead they migrate to the periphery where they remain permanently unresponsive or

anergic to further antigenic stimulation (Duty et al., 2008). The anergic B cells die rapidly in the

periphery mostly because of the lack of survival signals from the antigen specific T cells. The fourth

potential fate of a self-reactive B cell is that they ignore the antigenic stimulus which is too weak to

activate antigen receptor. Alternatively, some self-antigens may not be presented in the bone

marrow, and B cells specific for these self-antigens are left unaffected and enter the periphery.

Contrary to anergy, ignorance can be later counteracted if the concentration of the soluble antigen

suddenly increases so that the antigen receptor activation is possible. Thus, ignorant B cells may be

thought as a “seeds” of the autoimmune diseases. However, normally these low-affinity self reactive

B cells are not activated because of the lack of proper T cell help.

Only the B cells, which are not negatively selected in the bone marrow, have the potential to

further develop to antibody secreting mature B cells. However, still majority of the cells that enter

peripheral B cell pool are short lived and die before maturation is completed (Thomas et al., 2006).

These transitional B cells are dependent on the survival signals which they get at the lymphoid

follicles of peripheral lymphoid tissues. Continuous expression of B cell receptor seems to be

instrumental for B cell survival, since gene deletion experiments have shown that B cells lacking the

functional BCR complex are rapidly eliminated from the peripheral B cell pool (Kraus et al., 2004;

Torres et al., 1996). It seems that instead of antigen dependent activation of the receptor, antigen

independent signaling from the BCR is sufficient for survival (Monroe, 2006). In addition,

maturating B cells of the periphery also need other survival signals for their survival, such has

BAFF receptor activation by BAFF, a ligand belonging to TNF receptor family (Stadanlick and

Cancro, 2008).

30

B cell activation by T cells

When the maturating B cells first encounter their specific antigen, they have to be activated by

BCR-stimulation to generate effective immune responses. As in the case of immature B cells, the

antigenic stimulus leads primarily to deletion process. However, when antigen is delivered with co-

stimulatory signals derived by the activated T cells, the outcome is B cell activation instead of

deletion.

Figure 5. Crosstalk between B cell and a helper-T cell

Two events are necessary for antigen dependent B cell activation. Firstly, antigen binding triggers

the B cell receptor activation, antigen internalization and antigen presentation as peptides on the

surface of MHCII molecule. Secondly, B cell receives activating CD40-ligand (CD40-L) and

cytokines from an antigen specific T cell. Abbreviations: BCR; B cell receptor, TCR; T cell

receptor, MHCII; major histocompatibility complex II.

Crosstalk between a B cell and a helper-T cell specific to the same antigen is mandatory for B cell

activation (Figure 5.). After the first encounter of an antigen, antigen specific T and B cells are

trapped in border between T-and B-cell zones in the spleen and lymph nodes, where antigen is

presented in the surface of antigen presenting cells (dendritic cells) (MacLennan et al., 1997). After

B cell T cell

CD40 CD40-L

MHCII TCR

BCR

Antigen

Cytokines

31

antigen binding by BCR, the antigen is internalized and presented as peptides on the surface of a B

cell in MHCII-antigen complex, which is recognized by a T cell specific to the same antigen.

Subsequently, the T cell delivers contact mediated signals, of which the most important are CD40-

ligand and cytokines (IL4, IL5 and IL-6), to the B cell leading to its activation (Van den Eertwegh

et al., 1993).

Affinity maturation in germinal centers

Figure 6. The germinal center reaction Antigen activated B cell differentiates into centroblasts

that undergo proliferation and somatic hypermutation (SHM) in the dark zone of the germinal center

(GC). Centroblasts then move in the light zone of the GC and differentiate into centrocytes which

have three fates: centrocytes with self-reactive or non-binding antigen receptors are deleted by

apoptosis, while high affinity centrocytes are selected for further differentiation to memory B cells

or antibody secreting plasma cells.

After B cell activation, some of the B cells directly differentiate to antibody secreting short lived

plasma cells, which are responsible for the primary IgM-mediated responses against an antigen.

This primary response offers the first line defense against the invading agent, but the immune

Antigen activated B-cell

BCR

SHM

Centroblast

Centrocyte

1. Self-reactive

2. High affinity → Selection

3. Non-binding

Memory B cell

Plasma cell

Ig

CD4+ T cell

FDC

Apoptosis

Apoptosis Dark zone

Light zone

32

response is further improved by the antibody affinity maturation and isotype switching, which

ensures that antibodies will have more effective binding capacity and can evoke different effector

functions.

The affinity maturation takes place in germinal centers, which are formed by rapidly proliferating

B cells termed centroblasts (MacLennan, 1994) (Figure 6.). The vigorous proliferation is associated

with a somatic hypermutation (SHM), a process which introduces non-random, single base point

mutations into IgV-gene regions that encode the antigen binding site. Centroblasts then differentiate

into centrocytes, which are subjected for selection on the basis of their antigen binding capacity

(Figure 6.). Most of the centrocytes have gained deleterious mutations on their Ig-genes leading to

expression of a low affinity or a non-functional antigen receptor. These centrocytes are eliminated

by apoptosis. In addition, the SHM may lead to generation of B cells with self-reactivity, and also

these cells are eliminated by apoptosis because of the lack of survival signals (CD40-ligand)

delivered by activated T cells. Only the centrocytes with a high affinity BCR are able to bind

antigen efficiently and present it to helper T cells and follicular dendritic cells (FDC) to gain the

appropriate signal for further differentiation to antibody secreting plasma cells or memory B cells.

Some of the B cells may go repeated rounds of proliferation, SHM and selection to further improve

the antigen binding capacity. A subset of centrocytes undergoes immunoglobulin glass –switch

recombination, which results in the production of different immunoglobulin subclasses by plasma

cells (MacLennan, 1994).

2.2.4 Life and death decisions of a B cell; regulation of B cell apoptosis

To maintain B cell homeostasis, the high generation rate of B cells must be counteracted by

elimination of deleterious or extraneous B cells. Mainly, these unwanted cells are deleted by

apoptosis during various check points of B cell maturation process (see B cell maturation).

Apoptosis plays an important role in the negative selection of bone marrow immature B cells and

GC B cells. In addition, excess B lymphocytes are eliminated by apoptosis after the antigenic

challenge. The fate of a B cell is regulated by both survival and death signals originating from

stromal cells of lymphoid organs and activated T cells. Here, I will focus on the regulation of B cell

apoptosis during the GC reaction, and the signaling through antigen-, Fas-, and CD40-receptors, the

main regulators of B cell apoptosis. Knockout models of Fas and CD40 receptors and their

downstream mediators are developed to study their role in GC B cell development (table 1.).

33

Table 1. Knockout models of receptors and their downstream mediators controlling the

germinal center reaction.

The knockout models included are B cell-specific, so that only B cells were deficient of the targeted

gene.

Receptor Knockout model Phenotype Publication

Fas/CD95 Fas (the

lymphoproliferation

mutation)

Lymphadenopathy,

accumulation of heavily

mutated memory B cells

Takahashi et al., 2001;

Watanabe-Fukunaga et

al., 1992

Fas (GC B cell-specific) Fatal lymphoproliferation Hao et al., 2008

FADD Increased splenic and lymph

node B cells

Imtiyaz et al., 2006

CD40 CD40 Defective GC formation and

class switching

Kawabe et al., 1994

c-FLIP Reduced numbers of peripheral

B cells that were sensitive to

Fas-induced apoptosis

Zhang et al., 2009

NF-κB c-Rel Defect in B cell proliferation Cheng et al., 2003

NF-κB RelA Defective production of IgG

and IgA antibodies

Doi et al., 1997

Apoptosis during GC reaction

GC centrocytes are especially prone to apoptosis to ensure the rapid elimination of B cells with

newly generated non-functional or non-binding antigen receptors. Only few of the maturating B

cells have improved binding capacity for antigen after SHM and the rest of the B cells undergo cell

death by apoptosis and are removed by macrophages which are found abundantly in the GC. The

GC B cells are susceptible for both spontaneous and receptor mediated apoptosis, and these

mechanisms act in concert to ensure effective selection of B cell clones during the GC reaction.

Isolated GC cells and also their malign counterparts die spontaneously in vitro, if they are not

rescued by survival signals provided by follicular dendritic cells or T – helper cells (Eray et al.,

2003; Goval et al., 2008; Liu et al., 1989).

The death receptor CD95/Fas is one of the major regulators of the GC B cell apoptosis (van Eijk

et al., 2001). The negative selection of GC B cells was severely disrupted in mice with mutated Fas

receptor (the lymphoproliferation mutation), resulting in lymphadenopathy and accumulation of

self-reactive B cells with somatically mutated surface IgG (Takahashi et al., 2001; Watanabe-

Fukunaga et al., 1992). This finding was repeated also in another Fas-deficient mouse model, in

which non-functional Fas was associated with the accumulation of autoreactive and low-affinity B

cell clones (Hao et al., 2008). It has been suggested that the GC B cells contain preformed CD95

34

DISC, in which the caspase-8 can be activated spontaneously without the involvement of the Fas-

ligand (Hennino et al., 2001). In this model, the death receptor activation is inhibited by anti-

apoptotic c-FLIP (cellular FLICE inhibitory protein) which can interfere with caspase-8 in the pre-

formed DISC (Scaffidi et al., 1999). Without survival signals provided by CD40- or antigen

receptors or tropic factors from stromal cells, c-FLIP is rapidly degraded from the DISC leading to

caspase-8 activation and apoptosis (Hennino et al., 2001; van Eijk et al., 2001). However, Fas-

ligand expression at the mRNA level has been detected in GC T cells pointing out that the Fas

receptor activation might also be dependent on Fas ligand expressed on T-cells (Kondo et al., 1997).

The expression of Fas receptor is negatively correlated with the expression of anti-apoptotic Bcl-2

protein, since Bcl-2 protein level has been shown to drastically decrease during the GC reaction to

enhance Fas-mediated apoptosis (Kondo and Yoshino, 2007).

There is some controversy concerning the role B cell receptor stimulation in the selection process

of GC B cells. In the model of GC mediated selection presented above, BCR-mediated signal is

mainly proposed as a survival signal involved in the positive selection. However, according to

another hypothesis, signaling through the BCR is involved in the negative selection of somatically

mutated centrocytes with self-reactivity. This model is supported by findings that isolated BC B

cells and also their malign counterparts are susceptible for apoptosis induced by BCR triggering

(Billian et al., 1997; Eray et al., 2003). However, in both of the above mentioned models, activating

signals from CD4+ T cells (CD40L and IL-4), are critically involved in the positive selection of

high affinity, self tolerant centrocytes.

CD40-CD40L interactions in the regulation of B cell survival

CD40 is a 50 kDa transmembrane protein, which is expressed on B lymphocytes, monocytes and

dendritic cells and in addition several non-hematological tissues. It is also expressed by malignant

cells originating from these cells, including B and T cell lymphomas, multiple myeloma and

Hodgkin`s disease (Dallman et al., 2003). The ligand for CD40 (CD40L, CD154) is a

transmembrane protein expressed transiently on activated CD4+ T lymphocytes (van Kooten and

Banchereau, 2000; Younes and Kadin, 2003). In addition, CD40L expression has been documented

in activated B cells, natural killer cells, monocytes, basofils and dendritic cells. Constitutive

expression of CD40L has been demonstrated in a variety of B cell malignancies, including follicular

lymphoma, mantle cell lymphoma, diffuse large cell B cell lymphoma and chronic lymphoid

leukemia (Younes and Kadin, 2003).

35

The main function of CD40 is the regulation of T-cell mediated B-cell activation during humoral

immune response. The CD40 stimulation promotes B cell proliferation, immunoglobulin production

and class switching, GC formation and establishment of B cell memory (van Kooten and

Banchereau, 2000; Younes and Kadin, 2003).

The importance of CD40-CD40L interactions in the regulation of humoral immunity are

demonstrated in vivo in a human immunodeficiency disease X-linked hyper IgM syndrome in which

CD40 signaling is severely impaired due to mutations in the gene encoding CD40L. The disease is

characterized by accumulation of IgM type antibodies because of impaired Ig class switching, and

defects in the formation of GC and development of memory B cells (Korthauer et al., 1993). A

similar phenotype is presented also in CD40L or CD40 deficient mouse models (Dallman et al.,

2003).

Several in vitro studies have demonstrated the pivotal role of CD40 in the regulation of GC B cell

apoptosis. Isolated GC B cells die rapidly in vitro, unless they are not rescued by CD40L expressed

transiently on activated CD4+ T cells (Hennino et al., 2001). In addition, CD40 can counteract both

Fas and BCR -mediated apoptosis in both isolated GC B cells and in their malign counterparts

(Choe et al., 2000; Cleary et al., 1995; Eeva et al., 2003; Eeva et al., 2007).

CD40 contains a cytoplasmic tail, which can activate several intracellular signaling pathways by

binding adaptor molecules called TNF receptor associated factors (TRAFs) (van Kooten and

Banchereau, 2000). The activation of TRAF2 has been suggested as mediator of NF-κB activation

after CD40 ligation (Rothe et al., 1995). The NF-κB family includes the members p50, p65 (relA)

and c-Rel (Berberich et al., 1994), which are involved in the transcriptional control of several anti-

apoptotic genes including c-FLIP (Kreuz et al., 2001), survivin (Granziero et al., 2001), Bcl-2

(Holder et al., 1993), Bcl-xL and Bfl-1 (Lee et al., 1999). NF-κB activation is at least partially

mediated by a decrease in the half life of IκB-α and IκB-β proteins, which sequester NF-κB in the

cytosol in an inactive form (Schauer et al., 1996). Besides NF-κB activation, CD40 triggering leads

to the activation of other signaling pathways, including p38, JNK, ERK, JAK-STAT and NF-AT

pathways, but their importance in the regulation CD40 mediated survival is currently not well

characterized (van Kooten and Banchereau, 2000).

2.2.5 GC as an origin of B cell malignancies

The microenvironment that enhance vigorous cell proliferation, the SHM process, and a low

threshold for acquired mutations makes the GC B cells susceptible to malign transformation. The

36

hallmark of a malign transformation is the unregulated accumulation of single clone cells due to

overwhelming cell proliferation or inaccurate elimination by apoptosis. GC derived B cell

malignancies are most often associated with an aberrant Ig gene rearrangement, in which a highly

expressed Ig gene is joined next to genes that regulate cell growth or apoptosis. For example,

majority of the follicular lymphoma (FL) cells carry the chromosomal translocation t14:18 that

positions the Ig gene next to the antiapoptotic Bcl-2 gene leading to decreased apoptosis and

accumulation of long lived cells (Klein and Dalla-Favera, 2008). In Burkitt lymphoma cells the Myc

oncogene, involved in the control of cell cycle, is placed under the control of Ig loci leading to

enhanced cell proliferation (Klein and Dalla-Favera, 2008). Aberrant class switch recombination

leading to dysregulated expression of Bcl-6 may be the transforming event in the development of

DLBCL (diffuce large cell B cell lymphoma) (Klein and Dalla-Favera, 2008). Constitutive

expression of Bcl-6 results in the maintenance of proliferative, DNA-damage tolerant centroblastic

phenotype with subjects the cell for further genetic abnormalities.

The FL cells grow in a follicular pattern, and the same cellular components are found in the FL

follicles as are present in the normal germinal centers. It has been suggested that also the same

cellular interactions between FDCs, helper T cells and macrophages occur similarly in normal GC

B cells and FL cells (Kuppers, 2004). The dependence of FL cells as well as normal GC cells on the

growth support provided by their microenvironment is supported by the fact that the cells are

difficult to culture in vitro without survival signals provided by FDCs or CD40 (Eray et al., 2003;

Goval et al., 2008). Recently, a large scale gene-expression profiling study has demonstrated that

the clinical behavior of FL is not only determined by properties of tumor cells themselves, but also

by the molecular features of non-malignant tumor-infiltrating immune cells (Dave et al., 2004).

Another gene expression study demonstrated that especially the activation of CD4+ T-helper cells

play an important role in rapidly transforming FL with poor prognosis (Glas et al., 2007). In the

same study, it was found that the gene-expression profiles from of dendritic cells have an impact on

the growth properties and clinical behavior in FL. Further studies are needed for the better

understanding how the microenvironment specifically interacts with tumor cells, since this

knowledge can be used in the future for the development of new immunological treatment strategies

against FL.

37

2.3. Monoclonal antibody therapy for cancer treatment

In monoclonal antibody (mAb) therapy for cancer, the antibody against a specific tumor antigen is

given to patients as single therapy or in combination to chemotherapeutics to induce depletion of

cancer cells. In addition to cancer therapy, mAbs that target pro-inflammatory cytokines and their

receptors are currently used in the treatment of autoimmune diseases, such as rheumatoid arthritis,

systemic lupus erythematosus (SLE) and vasculitis (Bayry et al., 2007). The advantage of the mAb

therapy is the specificity of mAbs against tumor cells without a general toxicity associated with

conventional chemotherapeutics.

The target antigens currently used in the cancer therapy have several different biological

functions, such as regulation of cell death or proliferation, or conduction of growth promoting

signals inside the cell. Ideally, the target antigen should be expressed only in tumor cells, or in the

cell type from which the tumor cells arise. The expression of the antigen should not decrease upon

repeated mAb treatments, and the antigen should not detach from the cell surface upon antigen

binding.

Since the introduction of the hybridoma technique by Kohler and Milstein in 1975 (Kohler and

Milstein, 2005), specific mAbs have been available as large amounts needed for the therapy.

Hybridoma cells are produced by fusing a mouse myeloma cell with an antigen specific mouse B

cell extracted from a mouse immunized with the desired antigen. The resulting hybridoma cell line

is immortal and produces large amounts of the specific antibody. The pioneer study of mAb therapy

was conducted in the beginning of the 1980 century, when Nadler et. al applied intravenous mAb

against the tumor cell antigen to the lymphoma patient (Nadler et al., 1980). Although this pioneer

study showed no clinical response, it proofed for the first time that mAbs bind specifically to the

tumor cell antigen in vivo, and the therapy was not associated with a severe toxicity. In 1982, Levy

et al used an anti-idiotype antibody, an antibody specific to variable region of the B cell surface

immunoglobulin, to treat lymphoma patient (Miller et al., 1982). This clinical study showed that

mAb therapy can be directed specifically to lymphoma cells, since all lymphoma cells arise from a

single transformed cell and thus have uniform immunoglobulin variable regions (idiotype). The

lymphoma patient treated with anti-idiotype Ab had a complete response to the therapy showing the

curative potential of the treatment.

The first clinical studies of antibody therapy were conducted by using mouse antibodies, which

appeared to have several shortages in the clinical settings. The mouse antibodies are highly

immunogenic and thus induce the production of human-antimouse antibodies (HAMAs) as a result

38

of repeated dosing to the patients (Frodin et al., 1992). The mouse antibodies were associated with

mild adverse effects and the production of HAMAs resulted in too rapid clearance of the mAbs

from the circulation after their administration (Dillman et al., 1986). Later, the problems associated

with mouse antibodies were circumvented by the development of human-mouse chimeric antibodies

in which the Fab part of the immunoglobulin is mouse-derived and the Fc part of the antibody is

human specific (Figure 7.). More lately, the development of gene technology has enabled the

development of humanized monoclonal antibodies, in which only the six antigen binding

polypeptide loops (complementary determining regions (CDRs) are of non-human origin. For

example, the anti-CD52 antibody alemtuzumab, used for the treatment of chronic lymphocytic

leukemia, is produced by inserting mouse derived CDRs specific for CD52 into antibody

“framework” containing heavy and light chains of human origin (Crowe et al., 1992). Subsequently,

the humanized antibody, containing only CDRs of non-human origin, is expressed on immortalized

mouse cell line.

More recently, new technologies using phage display or transgenic mouse have emerged for the

generation of fully humanized antibodies (Lonberg, 2008). To date, two fully humanized

antibodies, adalimumab and panitumumab, are in the clinical use, and several antibodies in the

clinical trials. The anti-TNFα antibody adalimumab, used for the treatment of rheumatoid arthritis,

is produced by phage display technology. The antibody against epidermal growth factor receptor

(EGFR), panitumumab, is produced using “humanized” transgenic mice, which are able to produce

human immunoglobulin heavy and light chains after immunization. The antibody producing B cell

clone specific for EGFR is extracted from the transgenic mouse, and immortalized using Chinese

ovarian cell line.

The humanized and chimeric antibodies contain a human specific Fc part which can activate

leukocytes and complement system when applied to the human body. Indeed, antigen-dependent

cellular cytotoxicity (ADCC) and complement mediated cytotoxicity (CDC) are centrally involved

in the mAb-mediated tumor cell clearance in addition to the direct antigen specific growth-

inhibiting or apoptosis-inducing effect (Villamor et al., 2003) (Figure 8.). ADCC is mediated by

nonspecific cytotoxic cells such as NK cells, monocytes and macrophages which can recognize the

Fc part of a tumor-bound antibody by their Fcγ-receptors. The activation of Fcγ-receptors receptor

induces the secretion of lytic enzymes by the cytotoxic cells which results in the target cell lysis.

Complement dependent cytotoxicity is induced when antigen-antibody complex deposited on the

cell surface triggers the activation of classical complement pathway resulting in the formation of

39

large pores on the membrane of the target cell and cell lysis. In addition, ADCC and CDC can act

synergistically, so that the complement components deposited on the tumor cell surface facilitates

tumor cell clearance by recruiting phagocytic cells (complement dependent cellular cytotoxity,

CDCC).

To further improve immunotherapy, mAbs can be conjugated with toxins, chemotherapeutics and

radioisotopes, so that the resulting conjugated antibodies carry these cytotoxic agents specifically to

the tumor cells (Zinzani et al., 2008).

In the year 2009, six unconjugated humanized or chimeric monoclonal antibodies, along with one

radioisotope-conjugated antibody are approved for the clinical treatment of cancer in Finland (Table

2.). The success of currently used therapeutic antibodies has encouraged scientists to the search of

novel target antigens suitable for mAb therapy. Currently, several clinical trials are ongoing which

study the clinical effectiveness of novel mAbs.

Table 2. Therapeutic monoclonal antibodies approved for use in cancer therapy in Finland.

The antibodies listed below are used in the treatment of cancer according to Pharmaca Fennica

2009. In addition, there are several monoclonal antibodies used in the treatment of autoimmunity

diseases, including rheumatoid arthritis and crohn`s disease, which are not included in the table.

Abbreviations: EGFR; epidermal growth factor receptor, VEGF; vascular endothelial growth factor

Target Indication Reference

Unconjugated antibodies

Trastuzumab Her2/neu Breast cancer Guarneri et al., 2008

Rituximab CD20 Lymphoma Marcus and Hagenbeek, 2007

Cetuximab EGFR Colorectal cancer Raoul et al., 2009

Panitumumab EGFR Colorectal cancer Hecht et al., 2007

Bevacizumab VEGF Colorectal-, lung-,

kidney- and breast

cancer

Tol et al., 2009

Alemtuzumab CD52 Chronic

lymphocytic

leukemia

Schweighofer et al., 2009

Immunoconjugates

Ibritumoab tiuxetan

(Yttrium-90 conjugated)

CD20 Lymphoma Zinzani et al., 2008

40

2.4. Follicular lymphoma

Lymphoma is a cancer that originates from the lymphocytes of the immune system and presents as

a solid tumor of lymphoid cells. Traditionally, lymphomas are divided in two groups; hodgkin

lymphoma which was the first form of lymphoma ever described, and non-hodgkin lymphomas

(NHLs). More recently, lymphomas are grouped according to WHO classification that recognizes

over 40 different lymphoma types (Swerdlow et al., 2008).

2.4.1. Incidence and clinical characteristics

Non-Hodgin lymphomas (NHLs) represent a heterogeneous group of malignancies originating

mainly from B and T lymphocytes of the immune system. Follicular lymphoma (FL) is the second

most common NHL-subtype comprising 20% of the cases (Anderson et al., 1998). For yet unknown

reasons, the incidence of FL:s is increasing steadily (Chiu and Weisenburger, 2003). In Finland,

about 250 new cases of follicular lymphoma are diagnosed annually (Jyrkkiö et al., 2007).

The clinical course of FL is extremely variable. In some patients the disease proceed in an

indolent way over a period of many years whereas in other patients the disease transforms to diffuse

large cell B cell (DLBC) lymphoma with an aggressive course (Advani et al., 2004).

Patients with follicular lymphoma usually presents with an enlarged superficial lymph node.

Delays in diagnosis are not uncommon, since enlarged lymph nodes may be unnoticed for a

prolonged period of time. In some patients, lymph nodes grow in deep areas in the retroperitoneal,

iliac- or mesenteric regions. In those cases, patients may have atypical symptoms and tumor bulk

may be large at the time of diagnosis. Few patients have symptoms such as fatigue, weight loss and

night sweating at the time of diagnosis. The bone marrow is affected in about half of the cases.

Rarely, FL presents as extranodal tumor growth. (Salles, 2007)

2.4.2 Pathophysiology

FL originates from germinal center (GC) B cells of lymphoid tissue. GCs are the sites for antibody

affinity maturation, a process which enables the development effective immune response (Allen et

al., 2007). GC:s consist of rapidly dividing centroblasts undergoing somatic hypermutation and

centrocytes undergoing antigenic selection. Only the centrocytes with improved antigen binding

capacity escape from the selection and the rest of the centrocytes die by apoptosis. FL is thought to

arise from accumulation of inappropriately rescued long lived B cells due to resistance in apoptosis.

Classically, FL cells express anti-apoptotic Bcl-2 protein due to t14;18 translocation which places

41

the bcl-2 gene under the influence of the IgG gene enhancer (de Jong, 2005). However, the t14;18

translocation is not sufficient alone for lymphomagenesis but it enables the accumulation of

subsequent genetic lesions affecting proto-oncogenes and tumor-suppressor genes (Schwaenen et

al., 2009). Recently, gene expression studies have shown that tumor-infiltrating non-malignant

immune cells may influence the transformation risk and survival of FL patients (Dave et al., 2004).

The diagnosis of FL is based on the morphology and immunohistochemistry of an affected lymph

node. FL cells express the surface markers sIg (M, D, G, A), CD19, CD20, CD22 and CD79a which

can be detected by immunohistochemistry or by flow cytometry. In addition, germinal center B cell

antigens CD10 and Bcl-6 are expressed in FL cells (Marcus and Hagenbeek, 2007).

2.4.3 Current treatment of follicular lymphoma

The treatment and prognosis of FL has changed radically during the past years since the

introduction of monoclonal antibody therapies and stem cell transplantation. The standard

chemotherapies used at 1980-90s had no effect on overall survival rates and thus advanced stage FL

has been assumed as incurable disease. However, during the last years, the new therapies directed

against FL have markedly improved the life expectancy of FL patients.

The method of treatment is chosen on the basis of dissemination of a disease at the time of

diagnosis, which is stated by Ann Arbor classification (Table 3.) (Bendandi, 2008). In addition, the

age, symptoms and additional diseases of the patient contribute markedly to disease management.

The prognosis of FL is classified according to Follicular Lymphoma International Prognostic Index

(FLIPI) (Solal-Celigny et al., 2004) (Table 4.).

Table 3. The Ann Arbor classification

Stage Definition

I The cancer is located in a single region, usually one affected lymph node or one

extralymphatic organ

II Two or more affected lymph node regions, or local extralymphatic extension with lymph

nodes on the same side of the diaphgram

III Lymph node regions on both sides of the diaphgram, either alone or with local

extralymphatic extension

IV Diffuse involvement of one or more extralymphatic organs, including involvement of the

liver, bone marrow, or nodular involvement of the lungs

42

Table 4. Follicular lymphoma prognostic index (FLIPI)

Feature Good prognosis Adverse prognosis

Age ≤ 60 years > 60 years

Serum LDH level normal above normal

Ann Arbor stage I to II III to IV

Nodal sites involved ≤ 4 > 4

Hemoglobin level ≥ 120 g/l < 120 g/l

At most 20% of the patients have a regional disease without overall symptoms (Ann Arbor I-II) at

the time of diagnosis. In these cases radiotherapy of an affected lymph node is the most used

treatment modality. Radiotherapy can cure about half of the patients with stage I disease and quarter

of the patients with grade II FL (Wilder et al., 2001).

The majority of the patients have an advanced stage FL (Ann Arbor III-IV) at the time of

diagnosis. These patients are treated when disease related symptoms arise, FLIPI score is high or the

disease is morphologically aggressive (grade 3 which contains over 15 centroblasts per high-power

field) (Jyrkkiö et al., 2007). Elderly patients or patients with indolent disease and no symptoms

should first be managed by an active watch-and –wait approach because life expectancy of these

patients can not be improved by treatment (Ardeshna et al., 2003). Although a wide variety of

standard chemotherapeutics as single or combination regimen, with or without radiotherapy, have

been tested against FL, these treatment modalities have no effect on overall survival rates (Ardeshna

et al., 2003). However, at the last decade, the introduction of a monoclonal anti-CD20 antibody

rituximab has radically changed the treatment of B cell malignancies. Several large scale

randomized trials have proved that addition of rituximab to the standard chemotherapy significantly

improved clinical outcome (progression free survival and overall survival) in patients with

previously untreated advanced FL (Hiddemann et al., 2007). In addition, the use of rituximab as

maintenance therapy in relapsed FL responding to chemotherapy or immunotherapy has improved

progression free survival and overall survival rates (Marcus and Hagenbeek, 2007). The use of

rituximab maintenance therapy after first line therapy is currently being evaluated.

The rituximab therapy is usually well tolerated as compared to the standard chemotherapy which

is associated with considerable short and long term toxicity. Mild to moderate adverse effects, such

as hypotension, nausea and general weakness are relatively rarely reported after first infusion

(Kimby, 2005). Rituximab induces a rapid depletion of CD20 positive peripheral B cells, and the B

43

cell count remains low for the next 6 months (Kimby, 2005). Serum immunoglobulin levels remains

fairly stable, although a small reduction in serum IgM has been described during prolonged therapy.

Rituximab therapy has not associated with an increased risk to opportunist infections (Kimby,

2005), but further studies are needed to evaluate immunological adverse effects associated with the

long term use of rituximab.

In the future, the combination of anti-CD20 antibody with therapeutic radioisotopes may further

improve the survival of FL patients. Phase II study has demonstrated that 90-Ytrium-ibritumoab

(tiuxetan) either alone or as consolidation therapy followed by first line therapy improved response

rates in FL (Zinzani et al., 2008). Further studies are ongoing to establish the use of radiolabeled

anti-CD20 antibodies in FL treatment.

Myeloablative therapy followed by autologous stem cell transplantation (ASCT) and to a lesser

extend allogenic transplantation had improved long-term outcome of some patients with FL

(Hiddemann et al., 2007). These approaches are used not so commonly because of treatment related

toxicity and the risk of secondary malignancy, but may be considered in young patients with

relapsed FL with poor prognosis.

Rituximab- mechanism of action

Rituximab is a chimeric mouse/human monoclonal antibody targeted against human B cell surface

antigen CD20 (Figure 7.) (Reff et al., 1994). Several characteristics of rituximab make it excellent

tool for anticancer therapy. The expression pattern of CD20 on B cells is ideal, since it is expressed

on over 95% of B cell lymphomas (Mathas et al., 2000) and on normal B cells from the pre-B-cell

stage until the plasma cell differentiation (Reff et al., 1994). Thus, hematopoietic stem cells are left

unaffected by rituximab that enables the recovery of B cell pool after the therapy. In addition,

antibody secreting plasma cells do not express CD20, and thus the levels of antibodies are

maintained during the therapy. The humanized Fc part of the chimeric antibody enables the host

effector mechanism, which are important in rituximab induced cytotoxicity (see later). The

production of anti-chimeric antibodies is rare and not of clinical importance as compared to the

infusion of mouse mAbs which were associated with significant production of anti-mouse

antibodies (Kimby, 2005). Finally, the CD20 antigen is not internalized from the B cell surface

upon antibody binding (Reff et al., 1994).

Although the clinical results with rituximab have been promising, the exact mechanism how it

depletes B cells is controversial. Based on in vitro studies, rituximab can induce antibody-dependent

44

cell-mediated cytotoxicity (ADCC) (Reff et al., 1994), complement mediated cytotoxicity (CDC)

(Harjunpaa et al., 2000; Reff et al., 1994) and apoptosis (Mathas et al., 2000; Shan et al., 2000)

(Figure 8.). The relative contribution of these mechanisms to B cell depletion in vivo is not known,

and it is probable that synergistic action of these mechanisms is involved (Flieger et al., 2000). In

addition to direct induction of apoptosis, rituximab has been shown to sensitize lymphoma cells to

apoptosis induced by chemotherapeutics such as paclitaxel, fludarabine and cisplatin (Alas et al.,

2000; Alas et al., 2002; Jazirehi et al., 2003).

Figure 7. The structure of the rituximab antibody. The light- and heavy-chain variable regions are cloned from murine mAb 2B8 that specifically

recognizes CD20 molecule. The mouse derived genes encoding variable regions are linked with

genes encoding IgG1 heavy chain and kappa-light chain constant regions of human origin. After the

fusion of gene segments, the recombinant fusion antibody is produced in Chinese hamster ovarian

cells. (Reff et al., 1994). Abbreviations: FcγR; Fcγ-receptor

Variable region

Constant region

Binding region for FcγR

Light chain -

Heavy chain -

45

Figure 8. Mechanisms of rituximab action

The proposed mechanisms of action of rituximab include complement-mediated cytotoxicity,

antibody-dependent cell-mediated cytotoxicity and apoptosis. Abbreviations: B; B cell, FcγR; Fcγ-

receptor, MAC; membrane attack complex, NK; natural killer cell.

A) Complement-mediated cytotoxicity

B) Antibody-dependent cell-mediated cytotoxicity

C) Apoptosis

Rituximab

CD20

B B

Activation of the complement

MAC

Cytotoxic substances

FcγR CD20

Rituximab

CD20

Rituximab

NK NK B

B

B

46

3. AIMS OF THE STUDY

In general, the aim of the study was to examine the regulation of lymphoma cell apoptosis and

survival by cell surface receptors. The detailed knowledge of the apoptotic signaling pathways can

be used as an advantage in the development of new treatment modalities against B cell

malignancies.

Detailed aims were

1. to study the receptor-mediated regulation of B cell apoptosis during GC negative selection (I). To

get insight in the molecular mechanisms of B cell receptor – mediated apoptosis (I,III).

2. to study the molecular mechanisms of CD95/Fas –induced apoptosis (II, III). Could death

receptors serve as potential targets for cancer therapy?

3. to study the signaling pathways involved in the CD40-mediated protection against apoptosis (I,

II). Could CD40-CD40L interactions between lymphoma cells and the cells in their GC

microenvironment be responsible for resistance against cancer therapy? How the resistance

generated by anti-apoptotic function of CD40 can be counteracted?

4. to study the apoptotic signaling pathways induced by rituximab, and explore the means how

rituximab-induced cytotoxicity can be further enhanced (IV)

47

4. MATERIALS AND METHODS

4.1. Cell line and culturing

The studies presented in the original publications I-IV were performed in a human follicular

lymphoma (FL) cell line HF1A3. The cell line has been established from an enlarged lymph node of

a FL patient as described previously (Eray et al., 1994). The original patient sample from which the

cell line was established contained a nodular and diffuse centroblastic lymphoma according Kiel

classification (Eray et al., 2003). As a hallmark of a follicular lymphoma, HF1A3 cells contain the

t(14;18) translocation and overexpress the Bcl-2 protein. Based on the analysis of surface antigens

and responses against various physiological stimuli, the HF1A3 cells resemble germinal center

centrocytes undergoing negative selection. The cells were cultured as described in the original

publications I-IV.

4.2. Establishment of overexpressing cell lines

Transduction refers to the process where foreign DNA is introduced into cell by using a viral

vector. Commonly, the transduced gene coding for the particular protein is placed under the control

of a promoter that facilitates the transcription of the gene. In the publications III and IV, we used

transduced HF1A3 cell lines which overexpressed c-FLIP, Bcl-xL or dominat negative (DN) -

caspase-9 to study the signaling pathways essential for apoptosis. Transduction was performed by

using bicistronic lentivirus vector construct containing a gene of interest (DN-caspase-9, Bcl- xL,

FLIP-short or c-FLIP-long) followed by IRES (internal ribosomal entry site) and the GFP reporter

gene. Thus, the transduced cells always co-expressed the GFP reporter gene and the gene of interest.

After the transduction, the GFP-positive cells overexpressing the desired protein were sorted by

flow cytometry. Cells transduced with GFP reporter gene only were used as control cell line to

ensure that the cell transduction procedure in itself did not affect the viability of the cells, or made

them respond differently to apoptotic stimuli.

The establishment of DN-caspase-9 and Bcl-xL overexpressing cells has been described in detail

previously (Nuutinen et al., 2008). The establishment of c-FLIPlong/short overexpressing cells has

been described in the publication IV. Thus, only short overview of the cell transduction procedure is

given here.

48

Firstly, complementary DNA (cDNA) was synthesized by using the total mRNA extracted from

HF1A3 cells as a template. Thereafter, the sequences for caspase-9, Bcl-xL and c-FLIPshort / long were

amplificated by PCR using cDNA as template. Primer sequences for amplification contained SwaI-

sites which were designed to 5’ ends of the primers.

Dominant negative caspase-9 was generated by introducing a point mutation to catalytic cysteine,

Cys287Ser (TGT→AGT) by site-directed mutagenesis. This method uses oligonucleotide primer

containing the desired point mutation, and PCR to introduce the desired mutation to the gene of

interest.

The PCR fragments were isolated from agarose gel and inserted into a plasmid vector (pcr2.1-

TOPO-TA vector, Invitrogen, Carlsbad, CA). Subsequently, competent One Shot TOP10-E.coli

(Invitrogen) was transformed with the plasmid vector. Thereafter, the sequences were analysed to

avoid unwanted mutations. After sequencing inserts were cloned to SwaI site of the pWPI-IRES-

GFP (Tronolab, Switzerland) lentivirus-vector. In addition, pWPI-IRES-GFP lentivirus-vector was

produced and used for the establishment of control cell line. Subsequently, lentiviruses containing

the vector were produced following the protocol described earlier (Pellinen et al., 2004).

The HF1A3 cells were transduced with lentiviral vectors containing DN-caspase-9-IRES-GFP,

Bcl-xL-IRES-GFP, c-FLIPlong.-IRES-GFP and c-FLIPshort.-IRES-GFP. After three days, cells were

collected and cultured in fresh medium. The cultures were expanded and GFP positive cells were

enriched by using flow cytometer until over 95% of the cells were GFP positive. Overexpression of

Bcl-xL, DN-caspase-9, c-FLIPshort and c-FLIPlong was confirmed by Western blot.

4.3. Cell treatment experiments

In cell treatment experiments, apoptosis was induced by culturing cells in the presence of

antibodies directed against surface receptors. Physiologically, these receptors are activated by

binding of endogenous ligands that result in the clustering of adjacent receptors and activation of

intracellular signaling cascades. Similarly, antibodies used in the experiments cross-link adjacent

receptors resulting in their activation. The stimulating antibodies used in original publications are

introduced in table 5. Chemical inhibitors of signaling proteins were used to study importance of a

particular signaling pathway in apoptosis or survival. The inhibitors used are listed in the table 6.

The treatment of cells during the experiments is described in detail in the original publications.

49

Table 5. Antibodies used in cell treatment experiments. Abbreviations: κ; kappa, mAb;

monoclonal antibody, F(ab`)2; fragment antigen binding.

Antibody Type Antigen Publication

Anti-κ (anti-IgG) mAb, IgG1 κ-chain constant region of

human IgG

I, III

Anti-Fas (CH11) mAb, IgM Fas/ CD95 II, III, IV

AF.1.15 mAb, IgG4 CD40 I, II, IV

Rituximab chimeric, IgG1 CD20 IV

Anti-human

IgG(γ)

F(ab`)2 heavy chain constant region of

human IgG

IV

Table 6. Chemical inhibitors of signaling proteins

Inhibitor Target Publication

LY294002 phospahtidyl-inositol-3-kinase (PI3K) I

PD98059 mitogen-activated protein kinase kinase 1 (MEK1) I

Gö6850 protein kinase C (PKC) I

ZM336372 Raf I

PDTC NF- B II

z-IETD-fmk caspase-8 II

z-DEVD-fmk caspase-3 II, III

z-VAD-fmk caspases III, IV

4.4 Apoptosis detection

4.4.1 Quantification of fractional DNA content (I-IV)

Typical for apoptosis is the activation of endonucleases which cleave nuclear DNA at sites

between nucleosomes creating internucleosomal DNA fragments of about 200 base pairs and its

multiples. When DNA from apoptotic cells is analyzed by agarose gel electrophoresis, typical

“ladder” pattern of about 200 base pairs is detected. In the publications I-IV, flow cytometry and

the cell cycle analysis was used for the detection of cells with fragmented DNA. Before the

analysis, cells were fixed with ethanol and stained with a DNA binding dye propidium iodide.

50

As a result of DNA fragmentation, the amount of cells containing normal diploid chromosomal

content of DNA decreases, and a population of cells with decreased (hypodiploid) DNA content is

formed. The proportion of hypodiploid cells (sub-G1 fraction) can be quantified by flow cytometry.

4.4.2 Mitochondrial membrane potential depolarization (I-IV)

Mitochondrial membrane potential depolarization occurs as a result of changes in the

permeability of mitochondrial membranes during apoptosis. The depolarization can be analyzed

indirectly by using lipophilic cations which accumulate in the mitochondrial matrix driven by the

electrochemical gradient between the matrix and the intermembrane space. In publications II-IV,

loss of mitochondrial membrane potential ( Ψm) was assessed by using a cationic dye tetramethyl

rhodamine (TMRM, Molecular Probes) as previously described (Castedo et al., 2002). Cells lose

their capacity to retain TMRM stain when the electrochemical gradient between the matrix and

intermembrane space is collapsed. In the publication I, loss of mitochondrial membrane potential

( Ψm) was assessed by ApoAlert Mitochondrial Membrane Sensor Kit (Clontech Laboratories, Inc.,

California, USA). The kit contains a cationic dye MitoSensorTM

that fluoresces differently in

apoptotic and nonapoptotic cells, because of changes in mitochondrial membrane potential. After

staining, the cells were analyzed by flow cytometry. Only the early apoptotic cells with normal side-

and forward-scatter properties were included in the analysis.

4.4.3 Cytochrome c release (I-IV)

The analysis of cytochrome c release was based on the cell fractionation, which enabled the

separation of cell fraction containing mitochondria from the cytosol. The cell fractionation

procedure is described in detail in the publication I. After the fractionation, the amount of

cytochrome c in both mitochondrial and cytosolic fractions was detected by Western blotting.

Membranes were reprobed with an antibody against a mitochondrial protein cytochrome oxidase

subunit IV (COX-4) to ensure that the fractionation procedure was successful and mitochondrial

fractions contained equal amount of protein. Equal loading of cytosolic samples was assessed by

immunoblotting with polyclonal anti-actin antibodies.

4.4.4 Caspase activation (I-IV)

Commonly, more than one approach was used for the assay of caspase activity in original

publications. The activation of DEVD-specific caspases (i.e. caspase-3 and -7) was assessed by

51

employing fluorochrome substrate Ac-DEVD-AMC (Calbiochem, Ca, USA) which become

fluorescent upon cleavage by the caspase (publications I, II and IV). In addition to direct analysis of

caspase activity, activation of DEVD-specific caspases was assessed indirectly by testing whether

their specific inhibitor z-DEVD-fmk inhibited apoptosis (publications II and III). More specifically,

activation of caspase-3 was assayed by immunoblotting (publications II, III and IV). Upon

activation, full length caspase-3 is cleaved, and caspase activation can thus be detected by using an

antibody which recognizes both the cleavage product and the full length caspase-3.

The activation of caspase-8 was assayed indirectly by testing whether a chemical inhibitor z-

IETD-fmk, a specific inhibitor of caspase-8, inhibits apoptotic events (publication II). In addition,

cell lines expressing c-FLIP proteins, natural inhibitors of caspase-8 activation at the DISC, was

used to study the role of caspase-8 activation in apoptosis (publication III, IV). The activation of

caspase-8 was also assayed directly by immunoblotting as an increase of a cleavage product of

caspase-8 and a decrease of full length caspase-8. Recently, this method has been questioned since it

has been shown that merely dimerization, and not cleavage of caspase-8, is actually involved in its

activation (Pop et al., 2007). However, when caspase-8 activation was assessed in our experimental

model, we get similar results with immunoblotting and by inhibiting caspase-8 activation

chemically or by c-FLIP overexpression

The activation of caspase-9 was assayed by using cells stably overexpressing DN-caspase-9, in

which caspase-9 activity was blocked by single amino acid mutation at the catalytic site of the

protease (publications III, IV).

4.4.5 Cell membrane permeabilization III

During both necrotic and the late stages of apoptotic cell death, the cell membrane integrity is lost

and thus the cells can uptake stains such as propidium iodide, which can be detected by flow

cytometry. In the publication III, cell permeability assay was used as a cell viability analysis to

complement other assays of cell death. After staining with propidium iodide, cells were analyzed

with flow cytometry. The data were expressed as the proportion of living cells (the PI negative cells

with normal FCS properties) against the total cell count.

4.5. Western blotting

The samples containing equal amount of protein were run on SDS PAGE under reducing

conditions and transferred to nitrocellulose membranes. The membranes were pretreated with 3%

52

BSA buffer to reduce non-specific binding of antibodies. Following primary antibodies were used

for the immunoblotting: rabbit polyclonal anti-FLIPshort/long, rabbit polyclonal anti-caspase-3 H-277

and rabbit polyclonal anti-caspase-8 p20 (Santa Cruz Biotechnology, CA). Horseradish peroxidase-

conjugated goat anti-rabbit (Zymed Laboratories, CA) was used as a secondary antibody. ECL

chemiluminescence system (Amersham, Arlington Heights, IL) was used for detection.

4.6. Reverse transcriptional (RT)-polymerase chain reaction

In the publication II, C-FLIP mRNA expression was determined by quantitative reverse

transcription-polymerase chain reaction (RT-PCR). Total RNA extraction and PCR reaction were

performed as described in the publication II. The PCR products were analyzed on 2% agarose gel

and visualized by ethidium bromide staining.

4.7 Statistical analyses

In the publications I-II, student’s t-test was used to determine the statistical significance between

the selected groups. In the publication III, data were analyzed by repeated-measures

two-way ANOVA and Bonferroni post test. In the publication IV, Oneway ANOVA and Dunnett`s

post test was used for comparing selected group with other groups. Statistical analyses were

performed using GraphPad Prism version 4.0 for windows. p < 0.05 was considered significant.

53

5. RESULTS

5.1. The molecular mechanisms of B cell receptor-induced apoptosis (I, III)

The B cell receptor signaling has been a target of enthusiastic research during the past decades as

there are almost three thousand original articles published that concern BCR signaling. The BCR

has gained a lot of research interest because of its multiple and important functions in the B cell

development and adative immunity. As discussed earlier, the response to BCR-stimulation can be

proliferation, anergy or apoptosis depending on the maturational stage of a B cell and nature of

antigen binding (Niiro and Clark, 2002). This work studied the molecular mechanisms of BCR-

mediated apoptosis and how signals from the BCR regulate the fate of a germinal center B cell

coordinately with CD40 and Fas receptors.

5.1.1 The signal transduction pathways connected with B cell receptor-induced apoptosis

The signal transduction pathways involved in BCR-induced apoptosis are not yet clarified. Of the

major signaling pathways activated by BCR, the Ras-Raf-1-ERK and PLC-γ2 activation pathways

has been connected with apoptosis in some experimental models (Lee and Koretzky, 1998; Stang et

al., 2009) (Figure 9.). We used the specific inhibitors of Raf, ERK, PKC and p38 to study the

involvement of these kinases in BCR induced apoptosis (Figure 9.).

The Ras-Raf-1-ERK pathway was activated by BCR stimulation in HF1A3 cells, since we could

demonstrate a transient phosphorylation of ERK1/2 after BCR triggering analyzed by Western blot

(not shown). However, the selective inhibitors of Raf (ZM 336372) or ERK (PD 98059), showed

only slight reduction in the percentage of apoptotic cells after BCR triggering (Figure 10.). Pre-

treatment of cells with specific inhibitors of PKC (GÖ 6850) and p38 (SB 203580) showed only

minor effect on BCR-induced apoptosis (Figure 10.). Based on these results, we could not ascertain

the signal transduction pathway crucial for BCR-mediated apoptosis in HF1A3 cells.

54

Figure 9. The signal transduction pathways associated with B cell receptor-induced apoptosis. The specific inhibitors for the activation of Raf (ZM), PKC (GÖ), P38 (SB) and PI3K (LY) are

included in the picture.

5.1.2 The role of caspase-9 and mitochondrion in B cell receptor-induced apoptosis (I, III)

To get insight on the mechanism involved in BCR-induced apoptosis, we performed a careful

kinetic analysis of four sequential molecular events of apoptosis: activation of caspase-3,

fragmentation of nuclear DNA collapse of ΔΨm and release of cytochrome c (I). The activation of

caspase-3 was analyzed by a DEVD-specific protease assay and by Western blot analysis of the

caspase-3 cleavage product. The activation of caspase-3 started to increase 12 hours after BCR

triggering. Fragmentation of DNA, collapse of ΔΨm and release of cytochrome c followed similar

kinetics. BCR-induced apoptosis was dependent on the synthesis of new proteins, since the BCR-

induced DNA fragmentation and ΔΨm collapse were abrogated in the presence of the protein

synthesis inhibitor cycloheximide.

RAS

Raf

PLC

PKC Ca2+

PI3K

AKT

GSK3

Cyclin D NF-κB NFAT CREB

ERK JNK P38

BCR

ZM

BD

MAPKs

LY

SB

55

Figure 10. The inhibitors of the signaling pathways connected with BCR show no effect on

apoptosis. HF1A3 cells were cultured in media containing ERK inhibitor PD 98059 (15 μM), Raf

inhibitor ZM 336372 (10 μM), PKC inhibitor GÖ 6850 (500nM) or p38 inhibitor SB 203580 (10

μM) for one hour followed by stimulation with anti-IgG antibodies for additional 23 hours.

Subsequently, cells were harvested for propidium iodide staining followed by analysis of cells with

fragmented DNA by flow cytometry. The data are presented as mean + SD from three experiments.

The data are previously unpublished.

It is well established that caspase-9 is the apical caspase in the intrinsic pathway of apoptosis. To

explore whether caspase-9 is involved in the BCR-mediated apoptosis, we generated a DN-caspase-

9 overexpressing cell line in which caspase-9 was inactivated by a single amino acid mutation at the

catalytic domain of the protease (III). The cells transduced with the empty IRES-GFP-vector were

used as a vector control cells. The overexpression of DN-caspase-9 could effectively inhibit the

BCR-induced fragmentation of nuclear DNA (Figure 11b.) and cleavage of caspase-3. However,

BCR-induced Ψm collapse (Figure 11a.) and cell death measured by PI exclusion assay were only

partly inhibited by DN-caspase-9 expression. Interestingly, the BCR-induced cytochrome c release

was almost totally inhibited in DN-caspase-9 overexpressing cells, although the cytochrome c

release has been previously suggested to be upstream of caspase-9 activation in the apoptotic

cascade (III).

The release of cytochrome c and collapse of Ψm are regulated by a balance between pro-and anti-

apoptotic members of the Bcl-2 family (Kim et al., 2006; Willis et al., 2007). In consistent with this,

0

10

20

30

40

50

control PD anti-IgG anti-IgG

+ PD

0

10

20

30

40

50

control GÖ anti-IgG anti-IgG

+ GÖ

0

10

20

30

40

50

control ZM anti-IgG anti-IgG

+ ZM

0

10

20

30

40

50

control SB anti-IgG anti-IgG

+ SB

% a

popto

tic c

ells

% a

popto

tic c

ells

% a

popto

tic c

ells

% a

popto

tic c

ells

56

overexpression of Bcl-xL prevented both BCR-induced cytochrome c release and collapse of Ψm in

HF1A3 cells. In addition, the BCR-induced DNA fragmentation, caspase-3 activation and cell death

was blocked by Bcl-xL overexpression. Thus, we next asked if BCR-induced apoptosis was

associated with the activation of pro-apoptotic members of the Bcl-2 family. We could not detect

any changes in the expression level or intracellular location of pro-apoptotic Bim, Bik , Bad or Bax

proteins after BCR triggering, although these proteins have been associated with BCR-induced

apoptosis previously (Eldering et al., 2004; Jiang and Clark, 2001; Malissein et al., 2003; Takada et

al., 2006).

5.1.3 The caspase-8 activation pathway showed only marginal role in B cell receptor-induced

apoptosis

The previous work has demonstrated that cross-linking of BCR induces an apoptotic pathway

involving caspase-8 activation (Besnault et al., 2001). To study the role of caspase-8 activation

pathway in BCR-induced apoptosis in our experimental model we generated a cell line

overexpressing the long splice variant of cellular FLICE inhibitory protein, c-FLIPlong. Both the long

and short splicing variant of c-FLIP interfere with caspase-8 activation at the death inducing

signaling complex (DISC) (Krueger et al., 2001). HF1A3 cell were transduced with wild type c-

FLIPlong in IRES-GFP-lentivector or with IRES-GFP-lentivectror only (vector control).

The overexpression of c-FLIPlong only marginally reduced the appearance of hypodiploid cells or

Ψm collapse after BCR stimulation (Figure 11). The poor effect of c-FLIP overexpression on BCR-

induced apoptosis could not be explained by inadequate expression of c-FLIP, since Fas-induced

DNA fragmentation and Ψm collapse were almost completely blocked in the same c-FLIPlong

overexpressing cell line (Figure 11.). In concordance, the specific inhibitor of caspase-8 (z-IETD-

fmk) failed to inhibit BCR-mediated apoptosis while it effectively rescued cells from Fas-mediated

apoptosis.

Figure 11a-b. The effect of Bcl-xL, DN-caspase-9 and c-FLIPlong overexpression on receptor

mediated apoptosis. The cells overexpressing Bcl-xL, DN-caspase-9 or c-FLIPlong were cultured

with anti-IgG (3 μg/ml), rituximab (10 μg/ml) or anti-Fas (10 ng/ml) antibodies for 24 hours. Cells

transduced with IRES-GFP-lentivector only were used as vector control cells. A) After culture, the

cells were harvested for TMRM staining and flow cytometric analysis of cells with collapsed

mitochondrial membrane potential (TMRM low cells). B) Cells were harvested for propidium

iodide staining followed by analysis of cells with fractional DNA content by flow cytometry

(hypodiploid cells).

57

A)

B)

0

5

10

15

20

25

30

35

HF

1A

3-G

FP

HF

1A

3-B

cl-

XL

HF

1A

3-D

N-c

asp

-9

HF

1A

3-F

LIP

-L

HF

1A

3-G

FP

HF

1A

3-B

cl-

XL

HF

1A

3-D

N-c

asp

-9

HF

1A

3-F

LIP

-L

HF

1A

3-G

FP

HF

1A

3-B

cl-

XL

HF

1A

3-D

N-c

asp

-9

HF

1A

3-F

LIP

-L

HF

1A

3-G

FP

HF

1A

3-B

cl-

XL

HF

1A

3-D

N-c

asp

-9

HF

1A

3-F

LIP

-L

Control Anti-IgG Rituximab Anti-Fas

% c

ells

with

co

llap

sed

ΔΨ

m

0

10

20

30

40

50

60

HF1

A3-G

FP

HF1A

3-B

cl-X

L

HF

1A

3-D

N-c

asp

-9

HF1A

3-F

LIP

-L

HF

1A

3-G

FP

HF1A

3-B

cl-X

L

HF

1A

3-D

N-c

asp

-9

HF

1A

3-F

LIP

-L

HF

1A

3-G

FP

HF

1A

3-B

cl-X

L

HF1A

3-D

N-c

asp-9

HF

1A

3-F

LIP

-L

HF1A

3-G

FP

HF

1A

3-B

cl-X

L

HF1

A3

-DN

-cas

p-9

HF

1A

3-F

LIP

-L

Control Anti-IgG Rituximab Anti-Fas

% H

yp

od

iplo

id c

ells

58

5.2. The molecular mechanisms of Fas/CD95-induced apoptosis (II, III, IV)

Fas triggering in HF1A3 cells induced the cleavage and activation of caspase-8, which could be

demonstrated by Western blot as an increase of the 18 kDa cleavage product and a decrease of the

57 kDa full length caspase-8 (II). The specific inhibitor of caspase-8 activation, z-IETD-fmk,

completely inhibited Fas-induced apoptosis showing the necessity of caspase-8 activation for the

process (II). In line with this, the overexpression of both short and long splicing variants of c-FLIP,

which interfere with caspase-8 activation at the DISC and thus inhibit caspase-8 activation, blocked

Fas-induced apoptosis (Figure 11.) (IV). The activation of caspase-8 was followed by the activation

of caspase-3 and fragmentation of nuclear DNA (II). The activation of caspase-9, the initiator

caspase of the mitochondrial pathway, was unnecessary for Fas-induced apoptosis, since the

overexpression of catalytically inactive DN-caspase-9 showed no effect on cell death (Figure 11.)

(III). Fas-stimulation induced a profound release of cytochrome c and collapse of Ψm which could

be inhibited by Bcl-xL overexpression (Figures 11.) (III). However, the overexpression of Bcl-xL

showed no effect on Fas-induced activation of caspases -8 and -3, showing that the caspase

activation was independent on the mitochondrial breach (III). In addition, the overexpression of Bcl-

xL only partly inhibited Fas-induced apoptosis as measured by DNA fragmentation (Figure 11.) (III)

5.3. The molecular mechanisms of rituximab-induced apoptosis (IV)

Rituximab, the chimeric human-mouse mAb against the CD20 antigen, has been widely used in

the therapy against follicular lymphoma. Rituximab has been shown to deplete malign B cells by

three different and overlapping mechanisms: CDC, ADCC and apoptosis (Manches et al., 2003;

Reff et al., 1994; Shan et al., 1998; van der Kolk et al., 2002). Binding of rituximab to the CD20

surface antigen induces apoptosis by yet unknown mechanisms. Detailed knowledge of CD20-

induced signaling pathways to apoptosis may allow the development of more effective treatment

modalities against B cell malignancies.

Propensity of various B cell lineages to undergo CD20-mediated apoptosis is highly variable

(Chan et al., 2003; Manches et al., 2003). Sensitivity to rituximab-induced apoptosis has been linked

with a maturational stage of a lymphoma progenitor cell (Mattila and Meri, 2008). This finding was

also reproduced in our FL cell lines representing different maturational stages, as only one (HF1A3)

of three available cell lines responded to rituximab treatment by apoptosis. The other cell lines,

HF28 and H4.9, were resistant to rituximab-mediated apoptosis even when rituximab bound to

59

CD20 antigen was cross-linked by secondary antibody (anti-Fc -F(ab2)) (not shown). Although the

use of cross-linking secondary antibodies against rituximab has augmented apoptosis in some

experimental models (Stel et al., 2007; van der Kolk et al., 2002), in HF1A3 cells rituximab-induced

apoptosis could not be enhanced by hyper-cross-linking. Instead, in HF1A3 cells, F(ab2) fragments

directed against the Fc part of antigen receptor induced apoptosis itself.

The involvement of caspases in the rituximab-induced apoptosis has been controversially

discussed. To explore the necessity of the caspase-9 and caspase-8 for rituximab-induced apoptosis,

we developed cell lines in which the activity of these caspases was artificially blocked. The

involvement of caspase-9, the initiator caspase of the intrinsic apoptosis pathway, was studied by

using DN-caspase-9 overexpressing cells, in which caspase-9 was inactivated by a single amino

acid mutation at the catalytic domain of the protease. The HF1A3 cells overexpressing FLIPshort and

FLIPlong, natural inhibitor proteins of caspase-8 activation, were used to study the involvement of

caspase-8 activation in the rituximab-induced apoptosis. According to our results, the caspase-9

activation pathway was necessary for the rituximab-induced apoptosis, since blocking of the

caspase-9 activation inhibited all apoptotic events induced by rituximab, including the activation of

caspase-3, fragmentation of DNA (Figure 11b.), collapse of Ψm (Figure 11a) and release

cytochrome c (IV). In contrast, the caspase-8 activation pathway was dispensable for apoptosis

induced by rituximab, since the overexpression of FLIP proteins showed no effect on rituximab-

induced apoptosis although in the same time Fas-induced apoptosis was totally abrogated by c-FLIP

overexpression (Figure 11.).

Overexpression of Bcl-xL almost completely blocked the cytochrome c release and collapse of

Ψm, showing that these changes were regulated by Bcl-2 family proteins (Figures 11.) (IV). In

addition, rituximab-induced DNA fragmentation (Figure 11b.), collapse of Ψm (Figure 11a.),

caspase-3 activation and cell death were also abrogated in Bcl-xL overexpressing cells.

The kinetics and magnitude of rituximab-induced apoptosis were substantially slower compared

with Fas-induced apoptosis in the same cell line. After four days culture with rituximab, only about

half of the cells were apoptotic. The slow induction of apoptosis may be partly explained by the fact

that rituximab induces only a subpopulation of cells to undergo apoptosis at a time whereas the

remainder of the cells may still proliferate and repopulate the culture.

The simultaneous culture with anti-Fas and rituximab antibodies showed an additive effect on

apoptosis. The percentage of apoptotic cells increased from 15.9% +/- 1.7 (rituximab) to 38.3% +/-

3.0 (rituximab combined with anti-Fas).

60

5.4. Molecular mechanisms of CD40-mediated protection against apoptosis (I,

II and IV)

CD40 activation plays an important role in the regulation of B cell survival. In vivo, CD40

receptor is activated by its natural ligand CD40L (van Kooten and Banchereau, 2000; Younes and

Kadin, 2003). We studied CD40 signaling to reveal the signaling pathways involved in the anti-

apoptotic function of CD40. This knowledge could also be used in advantage in the study of

receptor-mediated apoptosis. Another goal was to get insight how B cell survival is regulated in the

germinal centers. The activation of the CD40 receptor was carried out by using an anti-CD40

antibody AF1.15. The specificity of the AF1.15 antibody for CD40 receptor has been studied by

immunoprecipitation (Pätäri A, 2000). CD40 stimulation could reverse apoptosis induced by Fas-

and B cell receptors (I, II), and in addition rituximab-induced apoptosis was inhibited by CD40

stimulation (Figure 12.) (IV).

0

10

20

30

40

50

60

Control Anti-IgG RTX Anti-Fas

% A

po

pto

tic c

ells

Without CD40 With CD40

Figure 12. The effect of CD40 stimulation of receptor-mediated apoptosis

HF1A3 cells were pre-treated two hours with anti-CD40 antibody or with medium only, followed

by culture with anti-IgG (3 μg/ml), rituximab (10 μg/ml) or anti-Fas (10 ng/ml) antibodies. After 24

hours culture, cells were harvested for propidium iodide staining and analysis of fractional DNA

content by flow cytometry. The data presented is collection from three separate experiments.

61

5.4.1 Kinetics and signaling requirements of CD40 mediated protection from B cell receptor-

mediated apoptosis (I)

As presented above, BCR-induced apoptosis in HF1A3 cells was manifested by caspase-3

activation, DNA fragmentation, cytochrome c release and collapse of Ψm. All these apoptotic

events were completely blocked by CD40 stimulation. Interestingly, CD40 stimulation was able to

protect from BCR-induced apoptosis up to 10 h from the induction of apoptosis. When CD40

stimulation was done 12 h or later after the induction of apoptosis, its protective effect was

gradually lost.

CD40 stimulation has been connected with the activation of the PI3-kinase, and mitogen activated

protein kinases (MAPKs) P38 and ERK (van Kooten and Banchereau, 2000). In addition, the PKC

activation pathway has been connected with B cell survival (Kronfeld et al., 2002). We used

specific inhibitors of these pathways to study their role in CD40-mediated survival.

Pre-treatment of cells with PI3-kinase inhibitor LY294002 had no effect on CD40-mediated

protection from BCR-induced apoptosis. Similar results were also observed with specific ERK

inhibitor PD 98059, Raf inhibitor ZM 336372, P38 inhibitor SB 203580 and PKC inhibitor Gö 6850

(I).

5.4.2 The CD40-induced protection against Fas-induced apoptosis was associated with a rapid

up-regulation of anti-apoptotic c-FLIP (II)

As discussed above, activation of caspase-8 was indispensable for the induction of Fas-induced

apoptosis. The Fas-induced cleavage and activation of caspase-8 was completely blocked by CD40

stimulation, since the p18 active cleavage product of caspase-8 could not be detected when cells

were pre-treated with CD40 before Fas stimulation. In addition, Fas-induced caspase-3 activation,

DNA fragmentation, cytochrome c release and Ψm collapse were blocked by CD40 stimulation.

The specific inhibitor of NF-κB activation, PDTC, was used to study the involvement of NF-κB in

the anti-apoptotic signaling of CD40. When cells were pre-treated with PDTC before stimulation

with Fas and CD40 antibodies, CD40 stimulation could no longer protect the cells from Fas-induced

apoptosis. In addition, in the presence of PDTC, CD40 stimulation could not inhibit the Fas-induced

caspase-8 activation.

Among other targets, transcription factor NF-κB regulates the expression of c-FLIP proteins,

which can interfere with caspase-8 activation at the DISC (Krueger et al., 2001). Thus, we next

studied if CD40 stimulation could induce c-FLIP proteins in our experimental model. CD40

62

stimulation induced a rapid upregulation of both long and short isoforms of c-FLIP, which could be

seen at the mRNA within 2 hours and protein level within 4 hours after CD40 stimulation (II). The

CD40-induced upregulation of c-FLIPshort/long mRNA and protein were completely blocked by

PDTC, showing that the up-regulation of c-FLIP was dependent on the activation of NF-κB (II).

63

6. DISCUSSION

6.1. FL cell lines as an experimental model for the study of receptor-mediated

apoptosis

The cell line HF1A3 provides an excellent model for the study of receptor-mediated apoptosis of

GC B cells, since the cells have retained the same immunophenotype and responses against

physiological stimuli as their natural counterparts (Eray et al., 2003). However, although the cell

line HF1A3 mimics germinal center B cells, the cells have genetic abnormalities such as Bcl-2

overexpression, which may hinder cell death. We used the HF1A3 cell line to study receptor-

mediated selection of GC B cells. The findings presented here would be more easily generalized to

normal GC B cells if we could replicate the results with primary GC B cells. However, the natural

GC B cells die rapidly in vitro without growth support provided by their microenvironment, which

makes it difficult to study apoptosis by using GC derived primary B cells. In addition, the cell

transfection studies could only be conducted with stable cell lines, since large amount of cells is

needed for the experimental procedure.

The cell line and the original patient sample have been carefully diagnosed according to REAL

classification as follicular center cell lymphoma, a subtype of a follicular lymphoma (Eray et al.,

2003). Thus, the cell line offers a versatile tool for the pre-clinical study of FL directed cancer

therapeutics. We used HF1A3 cell line for the study of molecular mechanisms involved in

rituximab-induced apoptosis. Unfortunately, only one of our three FL cell lines was susceptible for

rituximab-mediated apoptosis, and the other two cell lines (HF28 and HF4.9) were resistant. Thus,

our findings concerning the molecular mechanism of rituximab-induced apoptosis, as well as BCR

and Fas –induced apoptosis, need to be replicated in other B cell lines to generalize the results to all

B cells. The establishment of FL cell lines has been demanding, and still over 10 years after their

establishment, suitable human FL cell lines are not common.

64

6.2. Molecular mechanisms of receptor-mediated apoptosis

6.2.1 Fas-induced apoptosis proceed via caspase-8 activation pathway while B cell receptor-

and rituximab –induced apoptosis are dependent on mitochondrial pathway

As discussed earlier, two major apoptosis pathways have been classified; the death receptor

(extrinsic) and mitochondria dependent (intrinsic) pathways. The death receptor pathway relies on

the activation of the initiator caspase-8 at the DISC, followed by the activation of effector caspase-3

(Medema et al., 1997). In mitochondria dependent (intrinsic) apoptosis, various apoptosis-inducing

stress signals converge on mitochondria, resulting in the permeabilization of MOM, release of pro-

apoptotic IMS proteins and collapse of ΔΨm. The release of cytochrome c from IMS triggers the

activation of caspase-9, an initiator caspase of the intrinsic pathway (Li et al., 1997). Commonly,

the extrinsic pathway is engaged with the mitochondria-dependent pathway of apoptosis. In these

cases, caspase-8 cleaves pro-apoptotic BH3-only protein Bid, and the truncated Bid translocates to

mitochondria where it can promote MOMP and apoptosis (Li et al., 1998).

We used cell lines stably overexpressing Bcl-xL, DN-caspase-9 and c-FLIP to explore the

involvement the death receptor (extrinsic) or mitochondrial (intrinsic) pathways in receptor-

mediated apoptosis. As expected, Fas-induced apoptosis of HF1A3 cells showed typical

characteristics of extrinsic apoptosis. Firstly, Fas-induced apoptosis was completely prevented when

caspase-8 activation was blocked by a specific caspase-8 inhibitor z-IETD-fmk, or overexpression

of either short or long splicing variants of c-FLIP. Secondly, Fas-induced apoptosis proceeded

normally in cells overexpressing catalytically inactive form of caspase-9, which is an initiator

caspase of the intrinsic apoptosis. Finally, Bcl-xL overexpression showed only 50% decrease in the

amount of apoptotic (hypodiploid) cells, while Fas-induced cytochrome c release and collapse of

ΔΨm were markedly inhibited in Bcl-xL overexpressing cells. In line with this, we showed that

caspases were mainly activated upstream of mitochondrial breach in Fas-induced apoptosis.

However, although mitochondria seem not be involved in the activation of caspases during Fas-

induced apoptosis, the mitochondrial breach may still enhance the cell death process. Loss of

mitochondrial normal function leads to unbalanced Ca2+

release from the matrix, uncoupling of

oxidative phosporylation leading to ATP depletion and superoxide production and acidification of

cytoplasm due to anaerobic lactate production (Skommer et al., 2007). These events may ensure that

the cell death is effectively completed.

65

Apoptosis induced by BCR triggering and rituximab showed typical characteristics of intrinsic

apoptosis since in the both forms of apoptosis mitochondria were centrally involved.

Overexpression of Bcl-xL could inhibit cytochrome c release, collapse of ΔΨm and apoptosis

induced by BCR triggering and rituximab, showing the importance of mitochondrial permeability

change in apoptosis induced by BCR and rituximab. In addition, cells with catalytically inactive

caspase-9, an initiator caspase of the intrinsic pathway, were resistant to DNA fragmentation

induced by BCR and rituximab. Rituximab-induced apoptosis seemed even more dependent on

caspase-9 activity when compared to BCR-induced apoptosis, since rituximab-induced cell death

was almost completely prevented by DN-caspase-9 overexpression, while BCR-induced cell death

was only partially inhibited. Our results are in line with a previous study showing that DN-caspase-

9 block completely BCR-mediated PARP cleavage, DNA laddering and caspase activation, while

the cell death measured by PI exclusion decreased only about 50% (Herold et al., 2002). Based on

our findings, we suggest that BCR-mediated cytochrome c release and collapse of ΔΨm may be

induced in two phases, and only the second phase is mediated by caspase activation (see Figure 8. in

the publication III).

Finally, overexpression of c-FLIP did not affect either BCR- or rituximab-induced apoptosis

showing that the caspase-8 activation pathway was not involved. These findings are in variance with

previous studies, which demonstrated that caspase-8 activation pathway was needed for BCR – and

rituximab –induced apoptosis (Besnault et al., 2001; Stel et al., 2007). However, in these studies

anti-IgM and rituximab were hyper-cross-linked by secondary antibodies, and it may be that in these

circumstances extensive receptor cross-linking induces DISC-formation and caspase-8 activation.

6.2.2 Feedback regulation of mitochondria by caspase-9 during intrinsic apoptosis

It has been widely accepted that during intrinsic apoptosis, MOMP allows the release of

cytochrome c to the cytosol where it binds to Apaf-1 leading to formation of the caspase-9

activating complex called the apoptosome (Riedl and Salvesen, 2007). Thus, it was rather

unexpected to find out that both BCR- and rituximab -induced cytochrome c release was inhibited in

DN-caspase-9 cells showing that caspase-9 regulated the release of cytochrome c from the

mitochondria. This finding is in contrary to well documented view that cytochrome c release

controls the caspase-9 activation (Zou et al., 1999), and not vice versa. Moreover, rituximab-

induced collapse of Ψm was completely dependent and BCR-induced collapse of Ψm was

partially dependent on caspase-9 activation. Similarly, caspase-9 was demonstrated to control the

66

dexamethasone-induced collapse of Ψm in another FL cell line HF28 (Nuutinen et al., 2009). Thus,

the caspase-9 mediated feedback regulation of mitochondrial breach seems to be a common

phenomenon during the intrinsic apoptosis independent of the apoptotic insult, and not restricted to

only one cell line. In line with this, a recent study showed that caspases-3 and -7 are involved in the

collapse of Ψm and release of pro-apoptotic proteins, such as AIF, from the intermembrane space

during intinsic apoptosis (Lakhani et al., 2006). In the same study cytochrome c release into cytosol

was delayed in caspase-9 -/- mouse embryonic fibroblasts. The view that caspases can mediate Ψm

collapse is supported by a finding that the respiratory chain enzymes are cleaved and inactivated by

caspases during apoptosis (Ricci et al., 2004). Moreover, caspases can cleave and inactivate anti-

apoptotic members of the Bcl-2 family, including Mcl-1 and Bcl-xL, which can enhance the

cytochrome c release and Ψm collapse during apoptosis (Chen et al., 2007). Thus, it is possible that

also in our experimental model, caspase-9 and downstream caspases feed back to mitochondria by

cleaving respiratory chain enzymes and anti-apoptotic Bcl-2 proteins. This feed back signaling to

mitochondria may enhance apoptosis and ensure that the cell death process is effectively completed.

The conflicting results concerning the temporal ordering of cytochrome c release and caspase

activation may be explained by the hypothesis that cytochrome c is released in biphasic fashion.

Indeed, a previous study has shown that cytochrome c can be released from mitochondria at two

distinct stages, the early cytochrome c activation being caspase-independent and late stage massive

cyt c release being dependent on caspase activation (Chen et al., 2000). Thus, the first wave of

cytochrome c release may be caspase-independent event which results from incomplete MOMP.

This initial release of cytochrome c may be enough to activate the cascade of caspases which in turn

may induce the complete release of cytochrome c. Indeed, it has been demonstrated that caspases

can cleave and inactivate anti-apoptotic Bcl-2 family proteins and thus favor the release of

cytochrome c to the cytosol (Chen et al., 2007; Cheng et al., 1997).

As a conclusion, caspases may have a more prominent role in the control of mitochondrial

changes during intrinsic apoptosis than we have previously thought. We suggest that caspases may

feedback on mitochondria to enhance mitochondrial deterioration during intrinsic apoptosis.

67

6.2.3. Caspase-9 –mediated apoptosis was poorly inhibited by pan-caspase-inhibitor z-VAD-

fmk

Interestingly, we found that the widely used pan-caspase-inhibitor z-VAD-fmk showed no effect

on rituximab-induced apoptosis although apoptosis could be completely abrogated in the cells

overexpressing DN-caspase-9. Similarly, BCR-induced DNA fragmentation was inhibited by DN-

caspase-9 overexpression but not by z-VAD-fmk. In the same time, z-VAD-fmk blocked

completely DNA fragmentation and apoptosis induced by Fas showing that the inhibitor was

functional. Thus, it seems that the pan-caspase-inhibitor z-VAD-fmk may not effectively inhibit

caspase-9 activation while it is capable inhibiting other caspases such as caspase-8. In line with this,

over-expressed caspase-9 has been shown to process caspase-3 even in the presence of z-VAD-fmk

(Eldering et al., 2004). The poor effect of caspase-inhibitors against caspase-9 may be partly

explained by the fact that the binding of the inhibitor may be hindered because the caspase-9 is

buried in the apoptosome complex (Riedl and Salvesen, 2007). Based on these findings, the

activation of caspases should not be assessed solely by using caspase-inhibitors.

6.2.4. What happens before mitochondrial breach during intrinsic apoptosis?

We and others have demonstrated the several hours delay between the triggering of cell surface

receptors and mitochondrial changes during intrinsic apoptosis (Mackus et al., 2002). What happens

during this delay and how the signals originating from different apoptotic insults are integrated at

the level of mitochondria are central questions of apoptotic research. It is widely accepted that pro-

apoptotic Bcl-2 family proteins Bax and Bak are centrally involved in MOMP and apoptosis, since

cells deficient of both Bax and Bak have lost their susceptibility to MOMP and apoptosis (Wei et

al., 2001). It is also generally accepted, that the BH3-only activator proteins can either directly or

indirectly activate Bax and Bak (Kim et al., 2006; Willis et al., 2007). However, it is not known

which pro-apoptotic Bcl-2 proteins are involved and how they are activated in apoptosis induced by

BCR or CD20 stimulation.

Previous studies with mouse immature B cell line have shown that BCR-induced apoptosis involves

upregulation of proapoptotic Bcl-2 protein Bik and changes in Bad phosphorylation (Jiang and

Clark, 2001; Malissein et al., 2003). In the other study, JNK dependent phosphorylation and

activation of Bim was required for anti-IgM-induced apoptosis in mouse immature B cell line

(Takada et al., 2006). In addition, BCR- induced apoptosis and deletion of autoreactive B cells was

inhibited in Bim deficient mouse knock out model (Enders et al., 2003). However, the involvement

68

of Bim in human cell lines is not established as there may be differences in the involvement of

apoptotic Bcl-2 proteins in mouse and men. Currently, there are no studies exploring the roles of

Bcl-2 proteins during rituximab induced apoptosis.

The delay between BCR triggering and irreversible changes of apoptosis (caspase activation and

mitochondrial breach) was several hours. Based on our studies with a protein synthesis inhibitor

cycloheximide, we suggest that the delay enables a synthesis of new protein(s) which are

responsible for the mitochondrial breach and induction of apoptosis. We took advantage of 2-D

electrophoresis and mass spectroscopy to reveal if there was any BCR stimulation –induced changes

in the protein composition of mitochondria. Unfortunately, we could not find out new apoptosis-

regulating proteins by using this method (unpublished). Neither could we ascertain the early signal

transduction pathways originating from the BCR responsible for apoptosis.

Gene expression profiling has offered a new insight how apoptosis related gene expression can be

studied. In a recent study, CLL cells responding to IgM triggering by proliferation showed increased

expression of CD40, NF-κB and TRAF, which are associated with apoptosis inhibition (Guarini et

al., 2008). In the contrary, CLL cells showing apoptotic response failed to express these survival

genes, and it was concluded that the loss of survival signals leads to apoptosis in these cells.

Induction of pro-apoptotic genes could not be detected in the same experiment. In another gene-

expression study specific genes related with BCR-induced proliferation and apoptosis was studied

in Ramos cell line (Ollila and Vihinen, 2007). The study showed no increase in the genes associated

with intrinsic pathway of apoptosis, while there was an induction of some genes involved in the

death receptor pathway of apoptosis. Nevertheless, it has been demonstrated by previous studies that

the death receptor (extrinsic) pathway is not involved in BCR apoptosis (Berard et al., 1999;

Bouchon et al., 2000). Also in our studies, c-FLIP overexpression or the specific inhibitor of

caspase-8 z-IETD-fmk failed to inhibit BCR-induced apoptosis, showing that the caspase-8

activation pathway was not involved in BCR-induced apoptosis. As a conclusion, the identification

of the proteins synthesized during apoptosis remains a central goal for future research.

6.3. Molecular mechansims of CD40-mediated protection against apoptosis (I,

II and IV)

The main function of CD40 is the regulation of T-cell mediated B-cell activation during humoral

immune responses. CD40 activation promotes B cell proliferation, immunoglobulin production and

class switching, GC formation and establishment of B cell memory (van Kooten and Banchereau,

69

2000; Younes and Kadin, 2003). Most importantly, CD40 activation is involved in the regulation of

survival of both normal GC B cells and lymphoma cells. According to previous studies and our

findings, CD40 stimulation can reverse apoptosis induced by Fas-, BCR-, CD20- and TRAIL-

receptors ((Cleary et al., 1995; Mackus et al., 2002; Merino et al., 1995; Metkar et al., 1998; Travert

et al., 2008). In addition to TRAIL-induced apoptosis, apoptosis induced by dexamethasone and

Doxorubisine was attenuated by CD40-stimulation in FL cells (Nuutinen et al. 2009, accepted for

publication in Scand, J Immunol).

The signaling pathways responsible for anti-apoptotic function of CD40 are not completely

understood. Among the earliest signaling events following CD40 stimulation are the activation of

PI3K, MAPK P38 and Ras-Raf-1-ERK signaling pathways (Craxton et al., 1998; Gulbins et al.,

1996; Ren et al., 1994). However, these signaling pathways may not be the ones responsible for

survival signaling of CD40. Based on our results, the specific inhibitors of PI3K, P38, Raf-1 and

ERK could not inhibit the CD40-mediated rescue against BCR-induced apoptosis (I). Similar results

were repeated in a recent study, where inhibitors of PI3K, MAPK, P38 and JNK could not block

CD40-mediated rescue against TRAIL-induced apoptosis (Travert et al., 2008).

Accumulating evidence suggest that CD40-mediated survival is dependent on the activation of

NF-κB family members p50, p65 (relA) and c-Rel (Andjelic et al., 2000; Berberich et al., 1994;

Kreuz et al., 2001; Lee et al., 1999). NF-κB activation is at least partially mediated by a decrease in

the half life of IκB-a and IκB-b proteins, which sequester NF-κB in cytosol in inactive form

(Schauer et al., 1996). NF-κB is involved in the transcriptional control of several anti-apoptotic

genes including c-FLIP (Kreuz et al., 2001), survivin (Granziero et al., 2001), Bcl-2 (Holder et al.,

1993), Bcl-xL and Bfl-1 (Lee et al., 1999). According to previous studies and our results, it seems

that contribution of these different molecules to the CD40-mediated survival depends on the

apoptotic insult. Whereas c-FLIP proteins may have a dominant role in the CD40-mediated

protection against death receptor -mediated (extrinsic) apoptosis, the up-regulation of anti-apoptotic

members of the Bcl-2 family plays substantial role in the CD40-mediated survival against

mitochondria dependent (intrinsic) apoptosis (Kreuz et al., 2001; Lee et al., 1999) (Figure 13.).

Based on our findings, the CD40-mediated protection against death receptor -induced apoptosis

was at least partly mediated by NF-κB dependent up-regulation of c-FLIP proteins. Firstly, the Fas-

induced apoptosis could be completely blocked by c-FLIP expression. Secondly, CD40 stimulation

induced a rapid up-regulation of both c-FLIP proteins, and the c-FLIP up-regulation as well as the

anti-apoptotic function was completely dependent on the activation of NF-κB. In concordance, c-

70

FLIP proteins mediated also the anti-apoptotic function of CD40 against TRAIL-mediated apoptosis

(Travert et al., 2008). However, in addition to c-FLIP, also other NF-κB dependent proteins, such as

Bcl-xL, may be partly involved in the anti-apoptotic CD40 signaling against death receptor induced

apoptosis. It seems likely that CD40 stimulation induces a large array of protective molecules

which can interfere with different steps in apoptotic pathways (figure 13).

Figure 13. Molecular mechansims of CD40-mediated protection against apoptosis

CD40 stimulation induces the upregulation of c-FLIP, which interfere with caspase-8 activation

during extrinsic apoptosis. Moreover, CD40 up-regulates anti-apoptotic Bcl-2 proteins, including

Bcl-2 and Bcl-xL, which can stabilize mitochondrial membranes Overexpression of c-FLIP proteins

could not block mitochondria dependent apoptosis induced by rituximab or BCR-stimulation.

Instead, the overexpression of pro-apoptotic Bcl-2 protein Bcl-xL potently inhibited both BCR and

rituximab induced apoptosis. The pro-survival members of Bcl-2 family seem potential effectors of

CD40-mediated action for several reasons. Anti-apoptotic Bcl-2 family proteins inhibit cytochrome

c release and Ψm collapse by controlling mitochondrial membrane permeability (Kim et al., 2006;

Kuwana and Newmeyer, 2003; Scorrano and Korsmeyer, 2003). In concordance, the overexpression

Stress stimuli

Bax/Bak

Bc l-2 Bc l-XL

Mitochondria

Apoptosis

FLIP Apaf Cyt-c Cyt-c

Smac

HtrA2

IAPs

Death receptor

A) Extrinsic pathway B) Intrinsic pathway

FADD

Caspase-8

Bid

Caspase-3

Caspase-9

CD40 →

CD40 ↓

71

of Bcl-xL could inhibit BCR-induced Ψm collapse, cytochrome-c release and apoptosis in our

experimental model. In addition, CD40 stimulation could increase the expression of Bcl-xL both in

mRNA and protein level through NF-κB activation (Nuutinen et. al 2009, accepted for publication

in Scand J Immunol). Bcl-xL may not be the only pro-survival Bcl-2 protein involved in the anti-

apoptotic action of CD40, and other proteins including Mcl-1 or Bcl-2 might also be involved.

However, Bcl-2 alone seems not a very potent inhibitor of apoptosis, since the FL cells

overexpressing Bcl-2 were still susceptible for apoptosis induced by different stimuli.

.

6.4. A proposed model of the molecular interactions during the germinal

center negative selection

We studied the mechanisms of receptor-mediated apoptosis in the follicular lymphoma cell line

HF1A3. Based on the morphology, immunophenotype and propensity to antigen-induced apoptosis,

the cell line represents GC centrocytes undergoing negative selection (Eray et al., 2003). As in the

case of HF1A3 cell line, cancer cells often retain the same immunophenotype and responses against

physiological stimuli as their founder cells. Thus, the cell line offers a good model to study the

signaling requirements of GC B cells undergoing the negative selection. On the other hand, the

knowledge of the regulation of cell survival and death in GC can be used as an advantage for the

development of new treatment modalities against B cell derived tumors, since normal GC B cells

and lymphoma cells are similarly influenced by their GC microenvironment (Kuppers, 2004).

After the somatic hypermutaion of their Ig variable regions, centrocytes are subjected for selection

on the basis of their antigen binding capacity (Figure 6.) The centrocytes with a high-affinity BCR

are able to present antigen derived peptides in the contexts of MHC class II molecules for CD4+

helper T cells. Subsequently, the CD40 ligand is expressed transiently on the surface of helper T

cells, and delivered for the centrocyte specific for the same antigen. The activating signal through

both BCR and CD40 receptor selects the centrocyte for proliferation and differentiation in to

antibody secreting cells. On the other hand, centrocytes that have became autoreactive during the

SHM process are eliminated by the lack of proper T cell help, and die by apoptosis. Our findings

support the previous view that BCR stimulus may be involved in the negative selection of self-

reactive cells (Billian et al., 1997), since the cells die by apoptosis after BCR stimulation without

CD40 stimulus delivered within substantial time frame. The kinetics of BCR-mediated apoptosis

strengthens this hypothesis, since there was over 12 hours delay between receptor activation and

72

irreversible changes of apoptosis. In addition, CD40 stimulation was able to reverse apoptosis up to

10 hours after the BCR triggering. The time frame for CD40 rescue may be physiologically relevant

because specific T cell help may not be immediately available after the BCR stimulus. Before the

expression of CD40L is induced on the T cell surface, T cells have to be activated by interactions

between the T cell and the centrocyte.

Since SHM is a random process, only few of the centrocytes improve their antigen binding

capacity, whereas the majority of the centrocytes gain non-binding or non-functional receptors.

Based on studies with Fas deficient mouse models (Hao et al., 2008; Takahashi et al., 2001), Fas-

induced apoptosis is centrally involved in the selection process of GC B cells. Thus, the centrocytes

which have gained a low affinity BCR after the SHM may be eliminated by Fas-mediated apoptosis,

which ensures a rapid deletion of these unnecessary cells. It has been suggested that GC B cells

have a preformed CD95 DISC, in which caspase-8 activation is interfered by c-FLIP (Scaffidi et al.,

1999). Without survival signals provided by CD40- or antigen receptors or tropic factors from

stromal cells, c-FLIP is rapidly degraded from the DISC leading to caspase-8 activation and

apoptosis (Hennino et al., 2001; van Eijk et al., 2001). Thus, c-FLIP seems to be an important

regulator of the positive selection in GCs. However, the B cells with improved antigen binding

capacity which are rescued from Fas-induced apoptosis by c-FLIPlong expression may be still

susceptible for BCR-induced apoptosis, which is implicated in the negative selection of self-reactive

B cells (Billian et al., 1997). This suggestion is based on our findings that apoptosis induced by

BCR triggering was only marginally inhibited by c-FLIPlong overexpression.

As a conclusion, the BCR stimulus may be involved in both negative and positive selection of

centrocytes, and the T cell signals in the form of CD40 ligand discriminates between these fates. In

addition, death receptor mediated apoptosis may be necessary for the elimination of low affinity

centrocytes. The finding that Fas- and BCR -induced apoptosis use different signaling pathways

may be physiologically relevant, since it enables that Fas- and BCR-induced apoptosis can be

regulated independently, and targeted to different subsets of centrocytes. However, the hypothesis

presented here need to be replicated in normal GC B cells, since the FL cell line used may have got

genetic abnormalities during the long in vitro culture time, and thus not respond as original GC B

cells.

73

7. CONCLUSIONS; Therapeutic applications and future directions

Cell surface receptors which regulate the live and death of lymphocytes are important in the

regulation of normal B lymphocyte homeostasis. For example, extensive B cell proliferation must

be balanced by elimination of extraneous or potentially harmful cells during the germinal center

reaction. In addition to their importance in physiology, the cell surface receptors regulating cell

death serve as attractive targets for cancer therapy. This work was devoted on the study of four cell

surface receptors which are important in the regulation of B cell survival; B cell receptor, Fas/CD95

and CD20 are receptors mediating apoptosis, and CD40 which regulates anti-apoptotic response

against variety of apoptotic signals.

7.1. Cell surface receptors in the regulation of life and death decisions in the

germinal center

The first aim of the study was to explore the receptor mediated regulation of B cell survival

during the germinal center reaction. The germinal center B cells die spontaneously in vitro without

the growth support provided by GC microenvironment, and thus the mechanisms of receptor-

induced cell death is hard to study with the normal B cells. To circumvent this problem, we used a

follicular lymphoma cell line HF1A3 as an experimental model. Based on the morphology,

immunophenotype and propensity to antigen-induced apoptosis, the cell line represents GC

centrocytes undergoing negative selection (Eray et al., 2003). As in the case of HF1A3 cell line, the

malign cells often retain the same immunophenotype and responses against physiological stimuli as

their founder cells. Thus, the cell line offers a good model to study the signaling requirements of GC

B cells undergoing the negative selection.

Based on the existing literature and our findings of receptor mediated apoptosis, the following

model of GC selection process can be depicted (see also Figure 6.). After the SHM, the centrocyte

may have three different fates based on the antigen binding capacity of its newly formed antigen

receptor. The few centrocytes with improved antigen binding capacity are positively selected as

they get the CD40 survival signal from the antigen specific T cell due to effective antigen capture

and presentation to the helper T cell. On the other hand, the centrocytes in which SHM has resulted

non-functional or non-binding antigen receptor are negatively selected by the lack of T cell survival

74

signal CD40, and these cells are effectively eliminated by Fas-mediated apoptosis. The third fate of

a centrocyte is to have an autoreactive antigen receptor. In these cells the activation of antigen

receptor without a co-stimulatory CD40 signal leads to apoptosis. Thus, the signal from the B cell

receptor is involved both in the positive and negative selection of a centrocyte, and the T cell help in

the form of CD40 ligand discriminates between these fates.

The knowledge of the regulation of cell survival and death in GC can be used as an advantage for

the development of new treatment modalities against B cell derived tumors, as the normal GC B

cells and their malign counterparts are similarly influenced by their GC microenvironment

(Kuppers, 2004). For example, the survival signals provided by activated T cells may counteract the

cytotoxic effect of various chemotherapeutics, which may lead to resistance against cancer therapy.

7.2. CD40-CD40L interactions as potential targets for cancer therapy

As demonstrated by our FL cell lines in vitro, CD40 activation counteracts apoptosis induced by

cell surface receptors (Fas, TRAIL, CD20, BCR) and by chemotherapeutic agents (such as

dexamethasone and doxorubisin). Similarly, CD40 ligation may counteract in vivo the action of

various chemotherapeutics given to lymphoma patients, and thus generate resistance against cancer

therapy. This resistance problem might be circumvented by development of CD40 blocking agents,

which inhibit CD40-induced survival signaling. These CD40 blocking agents may function on the

receptor level, such as antibodies which bind to CD40L blocking its activation. Another approach is

to inhibit the downstream mediators of CD40 signaling. As demonstrated in this work, the most

important mediators of CD40 signaling include the up-regulation of c-FLIP and Bcl-xL proteins

which inhibit the death receptor-mediated and mitochondria-dependent apoptosis respectively.

Attractive approach might be blocking of these proteins by siRNA method, in which the mRNA of

these proteins is bound and cleaved specifically. Another approach might be the development of

Bcl-xL small molecule inhibitors. Indeed, several small molecule inhibitors against another survival

protein Bcl-2 are already developed and currently under clinical trials (Zeitlin et al., 2008).

However, it should be mentioned that in addition to its pro-survival function in several low grade

lymphomas, CD40 activation may be apoptotic signal in some more aggressive lymphomas and in

these circumstances CD40 agonistic antibody is more practical choice.

75

7.3. Novel approaches to enhance rituximab therapy

Until these days, FL has been suggested as incurable disease as the therapies used showed no

effect on the survival of FL patients. Introduction of rituximab has radically changed this view as its

use in combination with chemotherapy has improved the live time expectancy of FL patients.

However, still majority of the patients relapse after first line treatment. Thus efforts are being made

to develop more efficient therapy against FL.

One of the aims of this work was to study the molecular mechanism of rituximab induced

apoptosis. We demonstrated here the critical signaling events of rituximab-induced apoptosis in a

FL cell, which is currently one of the most important targets of rituximab therapy in the clinics. The

apoptosis induced by rituximab was dependent on the activation of caspase-9 and mitochondrial

breach, while the death receptor pathway was not involved. Thus, drugs designed to destabilize

mitochondrial membranes, such as BH3 mimetics or the antagonists of anti-apoptotic Bcl-2

proteins, may be used in the future to enhance rituximab-induced apoptosis.

We showed also evidence, that the simultaneous triggering of the death receptor pathway by Fas

enhanced significantly apoptosis induced by rituximab. The additive effect of simultaneous Fas and

rituximab treatment on apoptosis may be a result of simultaneous activation of both the extrinsic

(caspase-8-mediated) pathway by Fas and intrinsic (mitochondria dependent) pathway by rituximab.

The death receptors appear as ideal targets for cancer therapy, as their activation directly leads to

caspase-8 activation and initiation of the cell suicide program. The resistance against cancer therapy

is most often associated with mutation of the tumor suppressor gene p53, and this resistance could

be elegantly circumvented by death receptor mediated apoptosis, which can proceed independently

of p53 activation. However, the use of FAS-ligand or Fas agonistic antibodies is not possible in the

clinical settings due to the expression of Fas in hepatocytes and extreme hepatotoxicity

demonstrated in mouse models (reviewed in (Yonehara, 2002)). To circumvent this problem, small

molecules which can induce self-aggregation and ligand-independent activation of Fas receptor

complex have been introduced. Of these molecules, edelfosine (ET-18-OCH) and perifosine can

induce apoptosis in primary multiple myeloma cell line by recruitment of Fas receptor and

downstream apoptotic signaling molecules into lipid rafts, leading to activation of the extrinsic

pathway of apoptosis (Gajate and Mollinedo, 2007). Edelfosine is selectively taken up by cancer

cells, and its therapeutic potential against B cell malignancies is currently being addressed in a

phase II clinical study (Feller et al., 2005). In the future, it might be worth of studying the effect of

edelfosine – rituximab combination therapy against FL in the clinical settings.

76

In vitro studies have shown that rituximab-induced apoptosis was enhanced also by simultaneous

activation of the death receptor TRAIL (Daniel et al., 2007). Activation of death receptors TRAIL-

R1 or -R2 by their ligands or agonistic antibodies induces death receptor pathway in cancer cells of

different origin, including lymphoma cells (Younes and Kadin, 2003). TRAIL receptors are

expressed predominantly on tumor cells making them excellent targets for cancer therapy. Several

clinical trials investigating TRAIL as a potential anti-cancer drug are currently ongoing (Younes

and Kadin, 2003). The efficacy and safety of TRAIL and rituximab combination therapy for NHL

should be addressed in the clinical trial in the future.

Current trend in immunotherapy is the development of humanized antibodies which can be used in

the treatment of hematological malignancies and inflammatory diseases. The infusion of humanized

antibodies is associated with lesser side-effects caused by human anti-mouse-antibodies compared

with chimeric antibodies. The progress in gene technology has enabled the production of humanized

antibodies engineered to induce more efficient B cell depletion. For example, the humanized CD20

antibody ofatumumab is more effective than rituximab in promoting the activation of complement

dependent cell death (Pawluczkowycz et al., 2009). Another humanized CD20 antibody

ocrelizumab is associated with higher antibody dependent cytotoxicity as compared with rituximab

(Sikder and Friedberg, 2008). The tissue penetration of antibodies has been enhanced by the

development of “small-weight antibodies”, which consist of only single heavy- and light -chain

variable regions of human origin combined with engineered human IgG constant regions (Sikder

and Friedberg, 2008). Whether the novel humanized antibodies will supplant rituximab for

treatment of hematological malignancies is currently being assessed by clinical trials.

77

8. REFERENCES

Advani, R., Rosenberg, S.A., and Horning, S.J. (2004). Stage I and II follicular non-Hodgkin's

lymphoma: long-term follow-up of no initial therapy. J. Clin. Oncol. 22, 1454-1459.

Alas, S., Bonavida, B., and Emmanouilides, C. (2000). Potentiation of fludarabine cytotoxicity on

non-Hodgkin's lymphoma by pentoxifylline and rituximab. Anticancer Res. 20, 2961-2966.

Alas, S., Ng, C.P., and Bonavida, B. (2002). Rituximab modifies the cisplatin-mitochondrial

signaling pathway, resulting in apoptosis in cisplatin-resistant non-Hodgkin's lymphoma. Clin.

Cancer Res. 8, 836-845.

Allen, C.D., Okada, T., and Cyster, J.G. (2007). Germinal-center organization and cellular

dynamics. Immunity 27, 190-202.

Anderson, J.R., Armitage, J.O., and Weisenburger, D.D. (1998). Epidemiology of the non-

Hodgkin's lymphomas: distributions of the major subtypes differ by geographic locations. Non-

Hodgkin's Lymphoma Classification Project. Ann. Oncol. 9, 717-720.

Andjelic, S., Hsia, C., Suzuki, H., Kadowaki, T., Koyasu, S., and Liou, H.C. (2000).

Phosphatidylinositol 3-kinase and NF-κB/Rel are at the divergence of CD40-mediated proliferation

and survival pathways. J. Immunol. 165, 3860-7.

Ardeshna, K.M., Smith, P., Norton, A., Hancock, B.W., Hoskin, P.J., MacLennan, K.A., Marcus,

R.E., Jelliffe, A., Vaughan, G., Hudson, et al. (2003). Long-term effect of a watch and wait policy

versus immediate systemic treatment for asymptomatic advanced-stage non-Hodgkin lymphoma: a

randomised controlled trial. Lancet 362, 516-522.

Ashkenazi, A., and Dixit, V.M. (1999). Apoptosis control by death and decoy receptors. Curr. Opin.

Cell Biol. 11, 255-260.

Bayry, J., Lacroix-Desmazes, S., Kazatchkine, M.D., and Kaveri, S.V. (2007). Monoclonal antibody

and intravenous immunoglobulin therapy for rheumatic diseases: rationale and mechanisms of

action. Nat. Clin. Pract. Rheumatol. 3, 262-272.

Bendandi, M. (2008). Aiming at a curative strategy for follicular lymphoma. CA Cancer. J. Clin. 58,

305-317.

Berard, M., Mondiere, P., Casamayor-Palleja, M., Hennino, A., Bella, C., and Defrance, T. (1999).

Mitochondria connects the antigen receptor to effector caspases during B cell receptor-induced

apoptosis in normal human B cells. J. Immunol. 163, 4655-62.

Berberich, I., Shu, G.L., and Clark, E.A. (1994). Cross-linking CD40 on B cells rapidly activates

nuclear factor-κB. J. Immunol. 153, 4357-4366.

Besnault, L., Schrantz, N., Auffredou, M.T., Leca, G., Bourgeade, M.F., and Vazquez, A. (2001). B

cell receptor cross-linking triggers a caspase-8-dependent apoptotic pathway that is independent of

the death effector domain of Fas-associated death domain protein. J. Immunol. 167, 733-740.

78

Billian, G., Mondiere, P., Berard, M., Bella, C., and Defrance, T. (1997). Antigen receptor-induced

apoptosis of human germinal center B cells is targeted to a centrocytic subset. Eur. J. Immunol. 27,

405-14.

Boatright, K.M., Renatus, M., Scott, F.L., Sperandio, S., Shin, H., Pedersen, I.M., Ricci, J.E., Edris,

W.A., Sutherlin, D.P., Green, D.R., et al. (2003). A unified model for apical caspase activation.

Mol. Cell 11, 529-541.

Boatright, K.M., and Salvesen, G.S. (2003). Mechanisms of caspase activation. Curr. Opin. Cell

Biol. 15, 725-731.

Bouchon, A., Krammer, P.H., and Walczak, H. (2000). Critical role for mitochondria in B cell

receptor-mediated apoptosis. Eur. J. Immunol. 30, 69-77.

Busslinger, M. (2004). Transcriptional control of early B cell development. Annu. Rev. Immunol.

22, 55-79.

Campbell, K.S. (1999). Signal transduction from the B cell antigen-receptor. Curr. Opin. Immunol.

11, 256-264.

Castedo, M., Ferri, K., Roumier, T., Metivier, D., Zamzami, N., and Kroemer, G. (2002).

Quantitation of mitochondrial alterations associated with apoptosis. J. Immunol. Methods 265, 39-

47.

Chan, H.T., Hughes, D., French, R.R., Tutt, A.L., Walshe, C.A., Teeling, J.L., Glennie, M.J., and

Cragg, M.S. (2003). CD20-induced lymphoma cell death is independent of both caspases and its

redistribution into triton X-100 insoluble membrane rafts. Cancer Res. 63, 5480-5489.

Chen, M., Guerrero, A.D., Huang, L., Shabier, Z., Pan, M., Tan, T.H., and Wang, J. (2007).

Caspase-9-induced mitochondrial disruption through cleavage of anti-apoptotic BCL-2 family

members. J. Biol. Chem.

Chen, Q., Gong, B., and Almasan, A. (2000). Distinct stages of cytochrome c release from

mitochondria: evidence for a feedback amplification loop linking caspase activation to

mitochondrial dysfunction in genotoxic stress induced apoptosis. Cell Death Differ. 7, 227-233.

Chen, W., Wang, H.G., Srinivasula, S.M., Alnemri, E.S., and Cooper, N.R. (1999). B cell apoptosis

triggered by antigen receptor ligation proceeds via a novel caspase-dependent pathway. J. Immunol.

163, 2483-2491.

Chen, Y.R., Wang, X., Templeton, D., Davis, R.J., and Tan, T.H. (1996). The role of c-Jun N-

terminal kinase (JNK) in apoptosis induced by ultraviolet C and gamma radiation. Duration of JNK

activation may determine cell death and proliferation. J. Biol. Chem. 271, 31929-31936.

Cheng, E.H., Kirsch, D.G., Clem, R.J., Ravi, R., Kastan, M.B., Bedi, A., Ueno, K., and Hardwick,

J.M. (1997). Conversion of Bcl-2 to a Bax-like death effector by caspases. Science 278, 1966-1968.

79

Cheng, S., Hsia, C.Y., Leone, G., and Liou, H.C. (2003). Cyclin E and Bcl-xL cooperatively induce

cell cycle progression in c-Rel-/- B cells. Oncogene 22, 8472-8486.

Chiu, B.C., and Weisenburger, D.D. (2003). An update of the epidemiology of non-Hodgkin's

lymphoma. Clin. Lymphoma 4, 161-168.

Choe, J., Li, L., Zhang, X., Gregory, C.D., and Choi, Y.S. (2000). Distinct role of follicular

dendritic cells and T cells in the proliferation, differentiation, and apoptosis of a centroblast cell

line, L3055. J. Immunol. 164, 56-63.

Cleary, A.M., Fortune, S.M., Yellin, M.J., Chess, L., and Lederman, S. (1995). Opposing roles of

CD95 (Fas/APO-1) and CD40 in the death and rescue of human low density tonsillar B cells. J.

Immunol. 155, 3329-37.

Craxton, A., Shu, G., Graves, J.D., Saklatvala, J., Krebs, E.G., and Clark, E.A. (1998). p38 MAPK

is required for CD40-induced gene expression and proliferation in B lymphocytes. J. Immunol. 161,

3225-36.

Crowe, J.S., Hall, V.S., Smith, M.A., Cooper, H.J., and Tite, J.P. (1992). Humanized monoclonal

antibody CAMPATH-1H: myeloma cell expression of genomic constructs, nucleotide sequence of

cDNA constructs and comparison of effector mechanisms of myeloma and Chinese hamster ovary

cell-derived material. Clin. Exp. Immunol. 87, 105-110.

Custodio, J.B., Cardoso, C.M., Madeira, V.M., and Almeida, L.M. (2001). Mitochondrial

permeability transition induced by the anticancer drug etoposide. Toxicol. in. Vitro. 15, 265-270.

Dallman, C., Johnson, P.W., and Packham, G. (2003). Differential regulation of cell survival by

CD40. Apoptosis 8, 45-53.

D'Amelio, M., Tino, E., and Cecconi, F. (2008). The apoptosome: emerging insights and new

potential targets for drug design. Pharm. Res. 25, 740-751.

Daniel, D., Yang, B., Lawrence, D.A., Totpal, K., Balter, I., Lee, W.P., Gogineni, A., Cole, M.J.,

Yee, S.F., Ross, S., et al. (2007). Cooperation of the proapoptotic receptor agonist rhApo2L/TRAIL

with the CD20 antibody rituximab against non-Hodgkin lymphoma xenografts. Blood 110, 4037-

4046.

Dave, S.S., Wright, G., Tan, B., Rosenwald, A., Gascoyne, R.D., Chan, W.C., Fisher, R.I., Braziel,

R.M., Rimsza, L.M., Grogan, T.M., et al. (2004). Prediction of survival in follicular lymphoma

based on molecular features of tumor-infiltrating immune cells. N. Engl. J. Med. 351, 2159-2169.

de Jong, D. (2005). Molecular pathogenesis of follicular lymphoma: a cross talk of genetic and

immunologic factors. J. Clin. Oncol. 23, 6358-6363.

Dillman, R.O., Beauregard, J.C., Halpern, S.E., and Clutter, M. (1986). Toxicities and side effects

associated with intravenous infusions of murine monoclonal antibodies. J. Biol. Response Mod. 5,

73-84.

80

Doi, T.S., Takahashi, T., Taguchi, O., Azuma, T., and Obata, Y. (1997). NF-κB RelA-deficient

lymphocytes: normal development of T cells and B cells, impaired production of IgA and IgG1 and

reduced proliferative responses. J. Exp. Med. 185, 953-961.

Donepudi, M., Mac Sweeney, A., Briand, C., and Grutter, M.G. (2003). Insights into the regulatory

mechanism for caspase-8 activation. Mol. Cell 11, 543-549.

Duty, J.A., Szodoray, P., Zheng, N.Y., Koelsch, K.A., Zhang, Q., Swiatkowski, M., Mathias, M.,

Garman, L., Helms, C., Nakken, B., et al. (2008). Functional anergy in a subpopulation of naive B

cells from healthy humans that express autoreactive immunoglobulin receptors. J. Exp. Med. 206,

139-151.

Eeva, J., Postila, V., Mätto, M., Nuutinen, U., Ropponen, A., Eray, M., and Pelkonen, J. (2003).

Kinetics and signaling requirements of CD40-mediated protection from B cell receptor-induced

apoptosis. Eur. J. Immunol. 33, 2783-2791.

Eeva, J., Ropponen, A., Nuutinen, U., Eeva, S.T., Mätto, M., Eray, M., and Pelkonen, J. (2007). The

CD40-induced protection against CD95-mediated apoptosis is associated with a rapid upregulation

of anti-apoptotic c-FLIP. Mol. Immunol. 44, 1230-1237.

Eldering, E., Mackus, W.J., Derks, I.A., Evers, L.M., Beuling, E., Teeling, P., Lens, S.M., van Oers,

M.H., and van Lier, R.A. (2004). Apoptosis via the B cell antigen receptor requires Bax

translocation and involves mitochondrial depolarization, cytochrome C release, and caspase-9

activation. Eur. J. Immunol. 34, 1950-1960.

Enders, A., Bouillet, P., Puthalakath, H., Xu, Y., Tarlinton, D.M., and Strasser, A. (2003). Loss of

the pro-apoptotic BH3-only Bcl-2 family member Bim inhibits BCR stimulation-induced apoptosis

and deletion of autoreactive B cells. J. Exp. Med. 198, 1119-1126.

Eray, M., Postila, V., Eeva, J., Ripatti, A., Karjalainen-Lindsberg, M.L., Knuutila, S., Andersson,

L.C., and Pelkonen, J. (2003). Follicular lymphoma cell lines, an in vitro model for antigenic

selection and cytokine-mediated growth regulation of germinal centre B cells. Scand. J. Immunol.

57, 545-555.

Eray, M., Tuomikoski, T., Wu, H., Nordstrom, T., Andersson, L.C., Knuutila, S., and Kaartinen, M.

(1994). Cross-linking of surface IgG induces apoptosis in a bcl-2 expressing human follicular

lymphoma line of mature B cell phenotype. Int. Immunol. 6, 1817-27.

Feller, N., Jansen-van der Weide, M.C., van der Pol, M.A., Westra, G.A., Ossenkoppele, G.J., and

Schuurhuis, G.J. (2005). Purging of peripheral blood stem cell transplants in AML: a predictive

model based on minimal residual disease burden. Exp. Hematol. 33, 120-130.

Flieger, D., Renoth, S., Beier, I., Sauerbruch, T., and Schmidt-Wolf, I. (2000). Mechanism of

cytotoxicity induced by chimeric mouse human monoclonal antibody IDEC-C2B8 in CD20-

expressing lymphoma cell lines. Cell. Immunol. 204, 55-63.

Frodin, J.E., Lefvert, A.K., and Mellstedt, H. (1992). The clinical significance of HAMA in patients

treated with mouse monoclonal antibodies. Cell Biophys. 21, 153-165.

81

Gajate, C., and Mollinedo, F. (2007). Edelfosine and perifosine induce selective apoptosis in

multiple myeloma by recruitment of death receptors and downstream signaling molecules into lipid

rafts. Blood 109, 711-719.

Galluzzi, L., Maiuri, M.C., Vitale, I., Zischka, H., Castedo, M., Zitvogel, L., and Kroemer, G.

(2007). Cell death modalities: classification and pathophysiological implications. Cell Death Differ.

14, 1237-1243.

Garrido, C., Galluzzi, L., Brunet, M., Puig, P.E., Didelot, C., and Kroemer, G. (2006). Mechanisms

of cytochrome c release from mitochondria. Cell Death Differ. 13, 1423-1433.

Glas, A.M., Knoops, L., Delahaye, L., Kersten, M.J., Kibbelaar, R.E., Wessels, L.A., van Laar, R.,

van Krieken, J.H., Baars, J.W., Raemaekers, J., et al. (2007). Gene-expression and

immunohistochemical study of specific T-cell subsets and accessory cell types in the transformation

and prognosis of follicular lymphoma. J. Clin. Oncol. 25, 390-398.

Goldstein, J.C., Munoz-Pinedo, C., Ricci, J.E., Adams, S.R., Kelekar, A., Schuler, M., Tsien, R.Y.,

and Green, D.R. (2005). Cytochrome c is released in a single step during apoptosis. Cell Death

Differ. 12, 453-462.

Goval, J.J., Thielen, C., Bourguignon, C., Greimers, R., Dejardin, E., Choi, Y.S., Boniver, J., and de

Leval, L. (2008). The prevention of spontaneous apoptosis of follicular lymphoma B cells by a

follicular dendritic cell line: involvement of caspase-3, caspase-8 and c-FLIP. Haematologica 93,

1169-1177.

Granziero, L., Ghia, P., Circosta, P., Gottardi, D., Strola, G., Geuna, M., Montagna, L., Piccoli, P.,

Chilosi, M., and Caligaris-Cappio, F. (2001). Survivin is expressed on CD40 stimulation and

interfaces proliferation and apoptosis in B-cell chronic lymphocytic leukemia. Blood 97, 2777-

2783.

Graves, J.D., Draves, K.E., Craxton, A., Saklatvala, J., Krebs, E.G., and Clark, E.A. (1996).

Involvement of stress-activated protein kinase and p38 mitogen-activated protein kinase in mIgM-

induced apoptosis of human B lymphocytes. Proc. Natl. Acad. Sci. U. S. A. 93, 13814-13818.

Green, D.R., and Kroemer, G. (2004). The pathophysiology of mitochondrial cell death. Science

305, 626-629.

Guarini, A., Chiaretti, S., Tavolaro, S., Maggio, R., Peragine, N., Citarella, F., Ricciardi, M.R.,

Santangelo, S., Marinelli, M., De Propris, M.S., et al. (2008). BCR ligation induced by IgM

stimulation results in gene expression and functional changes only in IgV H unmutated chronic

lymphocytic leukemia (CLL) cells. Blood 112, 782-792.

Guarneri, V., Frassoldati, A., Bruzzi, P., D'Amico, R., Belfiglio, M., Molino, A., Bertetto, O.,

Cascinu, S., Cognetti, F., Di Leo, A., et al. (2008). Multicentric, randomized phase III trial of two

different adjuvant chemotherapy regimens plus three versus twelve months of trastuzumab in

patients with HER2- positive breast cancer (Short-HER Trial; NCT00629278). Clin. Breast Cancer.

8, 453-456.

82

Gulbins, E., Brenner, B., Schlottmann, K., Koppenhoefer, U., Linderkamp, O., Coggeshall, K.M.,

and Lang, F. (1996). Activation of the Ras signaling pathway by the CD40 receptor. J. Immunol.

157, 2844-50.

Hao, Z., Duncan, G.S., Chang, C.C., Elia, A., Fang, M., Wakeham, A., Okada, H., Calzascia, T.,

Jang, Y., You-Ten, A., et al. (2005). Specific ablation of the apoptotic functions of cytochrome C

reveals a differential requirement for cytochrome C and Apaf-1 in apoptosis. Cell 121, 579-591.

Hao, Z., Duncan, G.S., Seagal, J., Su, Y.W., Hong, C., Haight, J., Chen, N.J., Elia, A., Wakeham,

A., Li, W.Y., et al. (2008). Fas receptor expression in germinal-center B cells is essential for T and

B lymphocyte homeostasis. Immunity 29, 615-627.

Harjunpää, A., Junnikkala, S., and Meri, S. (2000). Rituximab (anti-CD20) therapy of B-cell

lymphomas: direct complement killing is superior to cellular effector mechanisms. Scand. J.

Immunol. 51, 634-641.

Hartley, S.B., Crosbie, J., Brink, R., Kantor, A.B., Basten, A., and Goodnow, C.C. (1991).

Elimination from peripheral lymphoid tissues of self-reactive B lymphocytes recognizing

membrane-bound antigens. Nature 353, 765-769.

Hashimoto, A., Okada, H., Jiang, A., Kurosaki, M., Greenberg, S., Clark, E.A., and Kurosaki, T.

(1998). Involvement of guanosine triphosphatases and phospholipase C-γ2 in extracellular signal-

regulated kinase, c-Jun NH2-terminal kinase, and p38 mitogen-activated protein kinase activation

by the B cell antigen receptor. J. Exp. Med. 188, 1287-1295.

Hecht, J.R., Patnaik, A., Berlin, J., Venook, A., Malik, I., Tchekmedyian, S., Navale, L., Amado,

R.G., and Meropol, N.J. (2007). Panitumumab monotherapy in patients with previously treated

metastatic colorectal cancer. Cancer 110, 980-988.

Hennino, A., Berard, M., Krammer, P.H., and Defrance, T. (2001). FLICE-inhibitory protein is a

key regulator of germinal center B cell apoptosis. J. Exp. Med. 193, 447-458.

Herold, M.J., Kuss, A.W., Kraus, C., and Berberich, I. (2002). Mitochondria-dependent caspase-9

activation is necessary for antigen receptor-mediated effector caspase activation and apoptosis in

WEHI 231 lymphoma cells. J. Immunol. 168, 3902-9.

Hiddemann, W., Buske, C., Dreyling, M., Weigert, O., Lenz, G., and Unterhalt, M. (2007). Current

management of follicular lymphomas. Br. J. Haematol. 136, 191-202.

Holder, M.J., Wang, H., Milner, A.E., Casamayor, M., Armitage, R., Spriggs, M.K., Fanslow, W.C.,

MacLennan, I.C., Gregory, C.D., and Gordon, J. (1993). Suppression of apoptosis in normal and

neoplastic human B lymphocytes by CD40 ligand is independent of Bc1-2 induction. Eur. J.

Immunol. 23, 2368-2371.

Hsu, Y.T., and Youle, R.J. (1998). Bax in murine thymus is a soluble monomeric protein that

displays differential detergent-induced conformations. J. Biol. Chem. 273, 10777-10783.

83

Imtiyaz, H.Z., Rosenberg, S., Zhang, Y., Rahman, Z.S., Hou, Y.J., Manser, T., and Zhang, J. (2006).

The Fas-associated death domain protein is required in apoptosis and TLR-induced proliferative

responses in B cells. J. Immunol. 176, 6852-6861.

Irmler, M., Thome, M., Hahne, M., Schneider, P., Hofmann, K., Steiner, V., Bodmer, J.L., Schroter,

M., Burns, K., Mattmann, C., et al. (1997). Inhibition of death receptor signals by cellular FLIP.

Nature 388, 190-5.

Irvine, R.A., Adachi, N., Shibata, D.K., Cassell, G.D., Yu, K., Karanjawala, Z.E., Hsieh, C.L., and

Lieber, M.R. (2005). Generation and characterization of endonuclease G null mice. Mol. Cell. Biol.

25, 294-302.

Jazirehi, A.R., Gan, X.H., De Vos, S., Emmanouilides, C., and Bonavida, B. (2003). Rituximab

(anti-CD20) selectively modifies Bcl-xL and apoptosis protease activating factor-1 (Apaf-1)

expression and sensitizes human non-Hodgkin's lymphoma B cell lines to paclitaxel-induced

apoptosis. Mol. Cancer. Ther. 2, 1183-1193.

Jiang, A., and Clark, E.A. (2001). Involvement of Bik, a proapoptotic member of the Bcl-2 family,

in surface IgM-mediated B cell apoptosis. J. Immunol. 166, 6025-6033.

Jyrkkiö, S., Jantunen, E., Keskinen, L., Kuittinen, O., Leppä, S., Vasala, K., Vornanen, M., and

Lehtinen, T. (2007). Follikulaarisen lymfooman hoito. Lääkärilehti 62, 1729-1733.

Kawabe, T., Naka, T., Yoshida, K., Tanaka, T., Fujiwara, H., Suematsu, S., Yoshida, N., Kishimoto,

T., and Kikutani, H. (1994). The immune responses in CD40-deficient mice: impaired

immunoglobulin class switching and germinal center formation. Immunity 1, 167-178.

Kerr, J.F., Wyllie, A.H., and Currie, A.R. (1972). Apoptosis: a basic biological phenomenon with

wide-ranging implications in tissue kinetics. Br. J. Cancer 26, 239-257.

Kim, H., Rafiuddin-Shah, M., Tu, H.C., Jeffers, J.R., Zambetti, G.P., Hsieh, J.J., and Cheng, E.H.

(2006). Hierarchical regulation of mitochondrion-dependent apoptosis by BCL-2 subfamilies. Nat.

Cell Biol. 8, 1348-1358.

Kimby, E. (2005). Tolerability and safety of rituximab (MabThera). Cancer Treat. Rev. 31, 456-

473.

Kischkel, F.C., Lawrence, D.A., Tinel, A., LeBlanc, H., Virmani, A., Schow, P., Gazdar, A., Blenis,

J., Arnott, D., and Ashkenazi, A. (2001). Death receptor recruitment of endogenous caspase-10 and

apoptosis initiation in the absence of caspase-8. J. Biol. Chem. 276, 46639-46646.

Klein, U., and Dalla-Favera, R. (2008). Germinal centres: role in B-cell physiology and malignancy.

Nat. Rev. Immunol. 8, 22-33.

Kluck, R.M., Bossy-Wetzel, E., Green, D.R., and Newmeyer, D.D. (1997). The release of

cytochrome c from mitochondria: a primary site for Bcl-2 regulation of apoptosis. Science 275,

1132-1136.

84

Kohler, G., and Milstein, C. (2005). Continuous cultures of fused cells secreting antibody of

predefined specificity. 1975. J. Immunol. 174, 2453-2455.

Koncz, G., Bodor, C., Kovesdi, D., Gati, R., and Sarmay, G. (2002). BCR mediated signal

transduction in immature and mature B cells. Immunol. Lett. 82, 41-49.

Kondo, E., Harashima, A., Takabatake, T., Takahashi, H., Matsuo, Y., Yoshino, T., Orita, K., and

Akagi, T. (2003). NF-ATc2 induces apoptosis in Burkitt's lymphoma cells through signaling via the

B cell antigen receptor. Eur. J. Immunol. 33, 1-11.

Kondo, E., and Yoshino, T. (2007). Expression of apoptosis regulators in germinal centers and

germinal center-derived B-cell lymphomas: insight into B-cell lymphomagenesis. Pathol. Int. 57,

391-397.

Kondo, E., Yoshino, T., Nishiuchi, R., Sakuma, I., Nishizaki, K., Kayagaki, N., Yagita, H., and

Akagi, T. (1997). Expression of Fas ligand mRNA in germinal centres of the human tonsil. J.

Pathol. 183, 75-79.

Korthauer, U., Graf, D., Mages, H.W., Briere, F., Padayachee, M., Malcolm, S., Ugazio, A.G.,

Notarangelo, L.D., Levinsky, R.J., and Kroczek, R.A. (1993). Defective expression of T-cell CD40

ligand causes X-linked immunodeficiency with hyper-IgM. Nature 361, 539-541.

Kraus, M., Alimzhanov, M.B., Rajewsky, N., and Rajewsky, K. (2004). Survival of resting mature

B lymphocytes depends on BCR signaling via the Igalpha/beta heterodimer. Cell 117, 787-800.

Kreuz, S., Siegmund, D., Scheurich, P., and Wajant, H. (2001). NF-κB inducers upregulate cFLIP, a

cycloheximide-sensitive inhibitor of death receptor signaling. Mol. Cell. Biol. 21, 3964-3973.

Kroemer, G., and Jäättelä, M. (2005). Lysosomes and autophagy in cell death control. Nat. Rev.

Cancer. 5, 886-897.

Kronfeld, I., Kazimirsky, G., Gelfand, E.W., and Brodie, C. (2002). NGF rescues human B

lymphocytes from anti-IgM induced apoptosis by activation of PKCzeta. Eur. J. Immunol. 32, 136-

43.

Krueger, A., Schmitz, I., Baumann, S., Krammer, P.H., and Kirchhoff, S. (2001). Cellular FLICE-

inhibitory protein splice variants inhibit different steps of caspase-8 activation at the CD95 death-

inducing signaling complex. J. Biol. Chem. 276, 20633-40.

Kuppers, R. (2004). Prognosis in follicular lymphoma--it's in the microenvironment. N. Engl. J.

Med. 351, 2152-2153.

Kuwana, T., Mackey, M.R., Perkins, G., Ellisman, M.H., Latterich, M., Schneiter, R., Green, D.R.,

and Newmeyer, D.D. (2002). Bid, Bax, and lipids cooperate to form supramolecular openings in the

outer mitochondrial membrane. Cell 111, 331-342.

Kuwana, T., and Newmeyer, D.D. (2003). Bcl-2-family proteins and the role of mitochondria in

apoptosis. Curr. Opin. Cell Biol. 15, 691-699.

85

Lakhani, S.A., Masud, A., Kuida, K., Porter, G.A.,Jr, Booth, C.J., Mehal, W.Z., Inayat, I., and

Flavell, R.A. (2006). Caspases 3 and 7: key mediators of mitochondrial events of apoptosis. Science

311, 847-851.

Lamkanfi, M., Festjens, N., Declercq, W., Vanden Berghe, T., and Vandenabeele, P. (2007).

Caspases in cell survival, proliferation and differentiation. Cell Death Differ. 14, 44-55.

Lee, H.H., Dadgostar, H., Cheng, Q., Shu, J., and Cheng, G. (1999). NF-κB-mediated up-regulation

of Bcl-x and Bfl-1/A1 is required for CD40 survival signaling in B lymphocytes. Proc. Natl. Acad.

Sci. U. S. A. 96, 9136-41.

Lee, J.R., and Koretzky, G.A. (1998). Extracellular signal-regulated kinase-2, but not c-Jun NH2-

terminal kinase, activation correlates with surface IgM-mediated apoptosis in the WEHI 231 B cell

line. J. Immunol. 161, 1637-44.

Li, H., Zhu, H., Xu, C.J., and Yuan, J. (1998). Cleavage of BID by caspase 8 mediates the

mitochondrial damage in the Fas pathway of apoptosis. Cell 94, 491-501.

Li, P., Nijhawan, D., Budihardjo, I., Srinivasula, S.M., Ahmad, M., Alnemri, E.S., and Wang, X.

(1997). Cytochrome c and dATP-dependent formation of Apaf-1/caspase-9 complex initiates an

apoptotic protease cascade. Cell 91, 479-89.

Liu, Y.J., Joshua, D.E., Williams, G.T., Smith, C.A., Gordon, J., and MacLennan, I.C. (1989).

Mechanism of antigen-driven selection in germinal centres. Nature 342, 929-31.

Lonberg, N. (2008). Fully human antibodies from transgenic mouse and phage display platforms.

Curr. Opin. Immunol. 20, 450-459.

Mackus, W.J., Lens, S.M., Medema, R.H., Kwakkenbos, M.J., Evers, L.M., Oers, M.H., Lier, R.A.,

and Eldering, E. (2002). Prevention of B cell antigen receptor-induced apoptosis by ligation of

CD40 occurs downstream of cell cycle regulation. Int. Immunol. 14, 973-82.

MacLennan, I.C. (1994). Germinal centers. Annu. Rev. Immunol. 12, 117-39.

MacLennan, I.C., Gulbranson-Judge, A., Toellner, K.M., Casamayor-Palleja, M., Chan, E., Sze,

D.M., Luther, S.A., and Orbea, H.A. (1997). The changing preference of T and B cells for partners

as T-dependent antibody responses develop. Immunol. Rev. 156, 53-66.

Malissein, E., Verdier, M., Ratinaud, M.H., and Troutaud, D. (2003). Changes in bad

phosphorylation are correlated with BCR-induced apoptosis of WEHI-231 immature B cells.

Biochimie 85, 733-740.

Manches, O., Lui, G., Chaperot, L., Gressin, R., Molens, J.P., Jacob, M.C., Sotto, J.J., Leroux, D.,

Bensa, J.C., and Plumas, J. (2003). In vitro mechanisms of action of rituximab on primary non-

Hodgkin lymphomas. Blood 101, 949-954.

Marcus, R., and Hagenbeek, A. (2007). The therapeutic use of rituximab in non-Hodgkin's

lymphoma. Eur. J. Haematol. Suppl. (67), 5-14.

86

Mathas, S., Rickers, A., Bommert, K., Dorken, B., and Mapara, M.Y. (2000). Anti-CD20- and B-

cell receptor-mediated apoptosis: evidence for shared intracellular signaling pathways. Cancer Res.

60, 7170-7176.

Mattila, A.M., and Meri, S. (2008). Responses to rituximab vary among follicular lymphoma B cells

of different maturation stages. Scand. J. Immunol. 68, 159-168.

Medema, J.P., Scaffidi, C., Kischkel, F.C., Shevchenko, A., Mann, M., Krammer, P.H., and Peter,

M.E. (1997). FLICE is activated by association with the CD95 death-inducing signaling complex

(DISC). EMBO J. 16, 2794-804.

Melamed, D., and Nemazee, D. (1997). Self-antigen does not accelerate immature B cell apoptosis,

but stimulates receptor editing as a consequence of developmental arrest. Proc. Natl. Acad. Sci. U.

S. A. 94, 9267-9272.

Merino, R., Grillot, D.A., Simonian, P.L., Muthukkumar, S., Fanslow, W.C., Bondada, S., and

Nunez, G. (1995). Modulation of anti-IgM-induced B cell apoptosis by Bcl-xL and CD40 in WEHI-

231 cells. Dissociation from cell cycle arrest and dependence on the avidity of the antibody-IgM

receptor interaction. J. Immunol. 155, 3830-3838.

Metkar, S.S., Naresh, K.N., Redkar, A.A., and Nadkarni, J.J. (1998). CD40-ligation-mediated

protection from apoptosis of a Fas-sensitive Hodgkin's-disease-derived cell line. Cancer Immunol.

Immunother. 47, 104-12.

Miller, R.A., Maloney, D.G., Warnke, R., and Levy, R. (1982). Treatment of B-cell lymphoma with

monoclonal anti-idiotype antibody. N. Engl. J. Med. 306, 517-522.

Monroe, J.G. (2006). ITAM-mediated tonic signalling through pre-BCR and BCR complexes. Nat.

Rev. Immunol. 6, 283-294.

Nadler, L.M., Stashenko, P., Hardy, R., Kaplan, W.D., Button, L.N., Kufe, D.W., Antman, K.H.,

and Schlossman, S.F. (1980). Serotherapy of a patient with a monoclonal antibody directed against

a human lymphoma-associated antigen. Cancer Res. 40, 3147-3154.

Nakagawa, T., Shimizu, S., Watanabe, T., Yamaguchi, O., Otsu, K., Yamagata, H., Inohara, H.,

Kubo, T., and Tsujimoto, Y. (2005). Cyclophilin D-dependent mitochondrial permeability transition

regulates some necrotic but not apoptotic cell death. Nature 434, 652-658.

Nemazee, D.A., and Burki, K. (1989). Clonal deletion of B lymphocytes in a transgenic mouse

bearing anti-MHC class I antibody genes. Nature 337, 562-566.

Niiro, H., and Clark, E.A. (2002). Regulation of B-cell fate by antigen-receptor signals. Nat. Rev.

Immunol. 2, 945-956.

Norvell, A., Mandik, L., and Monroe, J.G. (1995). Engagement of the antigen-receptor on immature

murine B lymphocytes results in death by apoptosis. J. Immunol. 154, 4404-4413.

87

Nuutinen, U., Ropponen, A., Suoranta, S., Eeva, J., Eray, M., Pellinen, R., Wahlfors, J., and

Pelkonen, J. (2009). Dexamethasone-induced apoptosis and up-regulation of Bim is dependent on

glycogen synthase kinase-3. Leuk. Res. 33, 1714-1717.

Nuutinen, U., Simelius, N., Ropponen, A., Eeva, J., Mätto, M., Eray, M., Pellinen, R., Wahlfors, J.,

and Pelkonen, J. (2008). PDTC enables type I TRAIL signaling in type II follicular lymphoma cells.

Leuk. Res.

Ollila, J., and Vihinen, M. (2007). Immunological systems biology: gene expression analysis of B-

cell development in Ramos B-cells. Mol. Immunol. 44, 3537-3551.

Pätäri A, K.M.,Anderson L. (2000). Abstracts of the SSi 31st annual meeting; (B14)

Characterization of a novel monoclonal antibody with a strong proliferating effect on human

peripheral blood B cells and an inhibitory effect on follicular lymphoma cells. Scand. J. Immunol.

52, 325.

Pawluczkowycz, A.W., Beurskens, F.J., Beum, P.V., Lindorfer, M.A., van de Winkel, J.G., Parren,

P.W., and Taylor, R.P. (2009). Binding of submaximal C1q promotes complement-dependent

cytotoxicity (CDC) of B cells opsonized with anti-CD20 mAbs ofatumumab (OFA) or rituximab

(RTX): considerably higher levels of CDC are induced by OFA than by RTX. J. Immunol. 183,

749-758.

Pellinen, R., Hakkarainen, T., Wahlfors, T., Tulimaki, K., Ketola, A., Tenhunen, A., Salonen, T.,

and Wahlfors, J. (2004). Cancer cells as targets for lentivirus-mediated gene transfer and gene

therapy. Int. J. Oncol. 25, 1753-1762.

Pop, C., Fitzgerald, P., Green, D.R., and Salvesen, G.S. (2007). Role of proteolysis in caspase-8

activation and stabilization. Biochemistry 46, 4398-4407.

Porter, A.G., and Urbano, A.G. (2006). Does apoptosis-inducing factor (AIF) have both life and

death functions in cells? Bioessays 28, 834-843.

Raoul, J.L., Van Laethem, J.L., Peeters, M., Brezault, C., Husseini, F., Cals, L., Nippgen, J., Loos,

A.H., and Rougier, P. (2009). Cetuximab in combination with irinotecan/5-fluorouracil/folinic acid

(FOLFIRI) in the initial treatment of metastatic colorectal cancer: a multicentre two-part phase I/II

study. BMC Cancer 9, 112.

Reff, M.E., Carner, K., Chambers, K.S., Chinn, P.C., Leonard, J.E., Raab, R., Newman, R.A.,

Hanna, N., and Anderson, D.R. (1994). Depletion of B cells in vivo by a chimeric mouse human

monoclonal antibody to CD20. Blood 83, 435-445.

Ren, C.L., Morio, T., Fu, S.M., and Geha, R.S. (1994). Signal transduction via CD40 involves

activation of lyn kinase and phosphatidylinositol-3-kinase, and phosphorylation of phospholipase C

γ 2. J. Exp. Med. 179, 673-80.

Reth, M., and Wienands, J. (1997). Initiation and processing of signals from the B cell antigen

receptor. Annu. Rev. Immunol. 15, 453-479.

88

Ricci, J.E., Munoz-Pinedo, C., Fitzgerald, P., Bailly-Maitre, B., Perkins, G.A., Yadava, N.,

Scheffler, I.E., Ellisman, M.H., and Green, D.R. (2004). Disruption of mitochondrial function

during apoptosis is mediated by caspase cleavage of the p75 subunit of complex I of the electron

transport chain. Cell 117, 773-786.

Richards, J.D., Dave, S.H., Chou, C.H., Mamchak, A.A., and DeFranco, A.L. (2001). Inhibition of

the MEK/ERK signaling pathway blocks a subset of B cell responses to antigen. J. Immunol. 166,

3855-3864.

Riedl, S.J., and Salvesen, G.S. (2007). The apoptosome: signalling platform of cell death. Nat. Rev.

Mol. Cell Biol. 8, 405-413.

Rothe, M., Sarma, V., Dixit, V.M., and Goeddel, D.V. (1995). TRAF2-mediated activation of NF-

κB by TNF receptor 2 and CD40. Science 269, 1424-1427.

Saito, M., Korsmeyer, S.J., and Schlesinger, P.H. (2000). BAX-dependent transport of cytochrome c

reconstituted in pure liposomes. Nat. Cell Biol. 2, 553-555.

Salles, G.A. (2007). Clinical features, prognosis and treatment of follicular lymphoma. Hematology

Am. Soc. Hematol. Educ. Program. 2007, 216-225.

Scaffidi, C., Fulda, S., Srinivasan, A., Friesen, C., Li, F., Tomaselli, K.J., Debatin, K.M., Krammer,

P.H., and Peter, M.E. (1998). Two CD95 (APO-1/Fas) signaling pathways. EMBO J. 17, 1675-87.

Scaffidi, C., Schmitz, I., Krammer, P.H., and Peter, M.E. (1999). The role of c-FLIP in modulation

of CD95-induced apoptosis. J. Biol. Chem. 274, 1541-8.

Schauer, S.L., Wang, Z., Sonenshein, G.E., and Rothstein, T.L. (1996). Maintenance of nuclear

factor-κB/Rel and c-myc expression during CD40 ligand rescue of WEHI 231 early B cells from

receptor-mediated apoptosis through modulation of IκB proteins. J. Immunol. 157, 81-6.

Schwaenen, C., Viardot, A., Berger, H., Barth, T.F., Bentink, S., Dohner, H., Enz, M., Feller, A.C.,

Hansmann, M.L., Hummel, M., et al. (2009). Microarray-based genomic profiling reveals novel

genomic aberrations in follicular lymphoma which associate with patient survival and gene

expression status. Genes Chromosomes Cancer 48, 39-54.

Schweighofer, C.D., Ritgen, M., Eichhorst, B.F., Busch, R., Abenhardt, W., Kneba, M., Hallek, M.,

and Wendtner, C.M. (2009). Consolidation with alemtuzumab improves progression-free survival in

patients with chronic lymphocytic leukaemia (CLL) in first remission: long-term follow-up of a

randomized phase III trial of the German CLL Study Group (GCLLSG). Br. J. Haematol. 144, 95-

98.

Scorrano, L., and Korsmeyer, S.J. (2003). Mechanisms of cytochrome c release by proapoptotic

BCL-2 family members. Biochem. Biophys. Res. Commun. 304, 437-444.

Shan, D., Ledbetter, J.A., and Press, O.W. (2000). Signaling events involved in anti-CD20-induced

apoptosis of malignant human B cells. Cancer Immunol. Immunother. 48, 673-683.

89

Shan, D., Ledbetter, J.A., and Press, O.W. (1998). Apoptosis of malignant human B cells by ligation

of CD20 with monoclonal antibodies. Blood 91, 1644-1652.

Sikder, M.A., and Friedberg, J.W. (2008). Beyond rituximab: the future of monoclonal antibodies in

B-cell non-Hodgkin lymphoma. Curr. Oncol. Rep. 10, 420-426.

Skommer, J., Wlodkowic, D., and Deptala, A. (2007). Larger than life: Mitochondria and the Bcl-2

family. Leuk. Res. 31, 277-286.

Solal-Celigny, P., Roy, P., Colombat, P., White, J., Armitage, J.O., Arranz-Saez, R., Au, W.Y.,

Bellei, M., Brice, P., Caballero, D., et al. (2004). Follicular lymphoma international prognostic

index. Blood 104, 1258-1265.

Stadanlick, J.E., and Cancro, M.P. (2008). BAFF and the plasticity of peripheral B cell tolerance.

Curr. Opin. Immunol. 20, 158-161.

Stang, S.L., Lopez-Campistrous, A., Song, X., Dower, N.A., Blumberg, P.M., Wender, P.A., and

Stone, J.C. (2009). A proapoptotic signaling pathway involving RasGRP, Erk, and Bim in B cells.

Exp. Hematol. 37, 122-134.

Stel, A.J., Ten Cate, B., Jacobs, S., Kok, J.W., Spierings, D.C., Dondorff, M., Helfrich, W., Kluin-

Nelemans, H.C., de Leij, L.F., Withoff, S., et al. (2007). Fas receptor clustering and involvement of

the death receptor pathway in rituximab-mediated apoptosis with concomitant sensitization of

lymphoma B cells to fas-induced apoptosis. J. Immunol. 178, 2287-2295.

Suzuki, M., Youle, R.J., and Tjandra, N. (2000). Structure of Bax: coregulation of dimer formation

and intracellular localization. Cell 103, 645-654.

Suzuki, Y., Imai, Y., Nakayama, H., Takahashi, K., Takio, K., and Takahashi, R. (2001). A serine

protease, HtrA2, is released from the mitochondria and interacts with XIAP, inducing cell death.

Mol. Cell 8, 613-621.

Swerdlow, S.H., Campo, E., Harris, N.L., Jaffe, E.S., Pileri, S.A., Stein, H., Thiele, J., and

Vardiman, J.W. (2008). In WHO Classification of Tumours of Haematopoietic and Lymphoid

Tissues, Fourth Edition , (Lyon: IARC)

Takada, E., Hata, K., and Mizuguchi, J. (2006). Requirement for JNK-dependent upregulation of

BimL in anti-IgM-induced apoptosis in murine B lymphoma cell lines WEHI-231 and CH31. Exp.

Cell Res. 312, 3728-3738.

Takahashi, Y., Ohta, H., and Takemori, T. (2001). Fas is required for clonal selection in germinal

centers and the subsequent establishment of the memory B cell repertoire. Immunity 14, 181-92.

Thomas, M.D., Srivastava, B., and Allman, D. (2006). Regulation of peripheral B cell maturation.

Cell. Immunol. 239, 92-102.

Thornberry, N.A., and Lazebnik, Y. (1998). Caspases: enemies within. Science 281, 1312-1316.

90

Tol, J., Koopman, M., Cats, A., Rodenburg, C.J., Creemers, G.J., Schrama, J.G., Erdkamp, F.L.,

Vos, A.H., van Groeningen, C.J., Sinnige, H.A., et al. (2009). Chemotherapy, bevacizumab, and

cetuximab in metastatic colorectal cancer. N. Engl. J. Med. 360, 563-572.

Torres, R.M., Flaswinkel, H., Reth, M., and Rajewsky, K. (1996). Aberrant B cell development and

immune response in mice with a compromised BCR complex. Science 272, 1804-1808.

Travert, M., Ame-Thomas, P., Pangault, C., Morizot, A., Micheau, O., Semana, G., Lamy, T., Fest,

T., Tarte, K., and Guillaudeux, T. (2008). CD40 ligand protects from TRAIL-induced apoptosis in

follicular lymphomas through NF-κB activation and up-regulation of c-FLIP and Bcl-xL. J.

Immunol. 181, 1001-1011.

Tsujimoto, Y., and Shimizu, S. (2007). Role of the mitochondrial membrane permeability transition

in cell death. Apoptosis 12, 835-840.

Van den Eertwegh, A.J., Noelle, R.J., Roy, M., Shepherd, D.M., Aruffo, A., Ledbetter, J.A.,

Boersma, W.J., and Claassen, E. (1993). In vivo CD40-gp39 interactions are essential for thymus-

dependent humoral immunity. I. In vivo expression of CD40 ligand, cytokines, and antibody

production delineates sites of cognate T-B cell interactions. J. Exp. Med. 178, 1555-1565.

van der Kolk, L.E., Evers, L.M., Omene, C., Lens, S.M., Lederman, S., van Lier, R.A., van Oers,

M.H., and Eldering, E. (2002). CD20-induced B cell death can bypass mitochondria and caspase

activation. Leukemia 16, 1735-1744.

van Eijk, M., Defrance, T., Hennino, A., and de Groot, C. (2001). Death-receptor contribution to the

germinal-center reaction. Trends Immunol. 22, 677-82.

van Kooten, C., and Banchereau, J. (2000). CD40-CD40 ligand. J. Leukoc. Biol. 67, 2-17.

Verhagen, A.M., Coulson, E.J., and Vaux, D.L. (2001). Inhibitor of apoptosis proteins and their

relatives: IAPs and other BIRPs. Genome Biol. 2, REVIEWS3009.

Verhagen, A.M., Ekert, P.G., Pakusch, M., Silke, J., Connolly, L.M., Reid, G.E., Moritz, R.L.,

Simpson, R.J., and Vaux, D.L. (2000). Identification of DIABLO, a mammalian protein that

promotes apoptosis by binding to and antagonizing IAP proteins. Cell 102, 43-53.

Villamor, N., Montserrat, E., and Colomer, D. (2003). Mechanism of action and resistance to

monoclonal antibody therapy. Semin. Oncol. 30, 424-433.

Watanabe-Fukunaga, R., Brannan, C.I., Copeland, N.G., Jenkins, N.A., and Nagata, S. (1992).

Lymphoproliferation disorder in mice explained by defects in Fas antigen that mediates apoptosis.

Nature 356, 314-317.

Wei, M.C., Zong, W.X., Cheng, E.H., Lindsten, T., Panoutsakopoulou, V., Ross, A.J., Roth, K.A.,

MacGregor, G.R., Thompson, C.B., and Korsmeyer, S.J. (2001). Proapoptotic BAX and BAK: a

requisite gateway to mitochondrial dysfunction and death. Science 292, 727-730.

91

Wilder, R.B., Jones, D., Tucker, S.L., Fuller, L.M., Ha, C.S., McLaughlin, P., Hess, M.A.,

Cabanillas, F., and Cox, J.D. (2001). Long-term results with radiotherapy for Stage I-II follicular

lymphomas. Int. J. Radiat. Oncol. Biol. Phys. 51, 1219-1227.

Willis, S.N., Fletcher, J.I., Kaufmann, T., van Delft, M.F., Chen, L., Czabotar, P.E., Ierino, H., Lee,

E.F., Fairlie, W.D., Bouillet, P., et al. (2007). Apoptosis initiated when BH3 ligands engage

multiple Bcl-2 homologs, not Bax or Bak. Science 315, 856-859.

Yang, J., Liu, X., Bhalla, K., Kim, C.N., Ibrado, A.M., Cai, J., Peng, T.I., Jones, D.P., and Wang, X.

(1997). Prevention of apoptosis by Bcl-2: release of cytochrome c from mitochondria blocked.

Science 275, 1129-1132.

Yin, Q., Park, H.H., Chung, J.Y., Lin, S.C., Lo, Y.C., da Graca, L.S., Jiang, X., and Wu, H. (2006).

Caspase-9 holoenzyme is a specific and optimal procaspase-3 processing machine. Mol. Cell 22,

259-268.

Yonehara, S. (2002). Death receptor Fas and autoimmune disease: from the original generation to

therapeutic application of agonistic anti-Fas monoclonal antibody. Cytokine Growth Factor Rev. 13,

393-402.

Youle, R.J., and Strasser, A. (2008). The BCL-2 protein family: opposing activities that mediate cell

death. Nat. Rev. Mol. Cell Biol. 9, 47-59.

Younes, A., and Kadin, M.E. (2003). Emerging applications of the tumor necrosis factor family of

ligands and receptors in cancer therapy. J. Clin. Oncol. 21, 3526-3534.

Zamzami, N., Susin, S.A., Marchetti, P., Hirsch, T., Gomez-Monterrey, I., Castedo, M., and

Kroemer, G. (1996). Mitochondrial control of nuclear apoptosis. J. Exp. Med. 183, 1533-1544.

Zeitlin, B.D., Zeitlin, I.J., and Nor, J.E. (2008). Expanding circle of inhibition: small-molecule

inhibitors of Bcl-2 as anticancer cell and antiangiogenic agents. J. Clin. Oncol. 26, 4180-4188.

Zhang, H., Rosenberg, S., Coffey, F.J., He, Y.W., Manser, T., Hardy, R.R., and Zhang, J. (2009). A

role for cFLIP in B cell proliferation and stress MAPK regulation. J. Immunol. 182, 207-215.

Zinzani, P.L., d'Amore, F., Bombardieri, E., Brammer, C., Codina, J.G., Illidge, T., Jurczak, W.,

Linkesch, W., Morschhauser, F., Vandenberghe, E., et al. (2008). Consensus conference:

implementing treatment recommendations on yttrium-90 immunotherapy in clinical practice - report

of a European workshop. Eur. J. Cancer 44, 366-373.

Zou, H., Li, Y., Liu, X., and Wang, X. (1999). An APAF-1.cytochrome c multimeric complex is a

functional apoptosome that activates procaspase-9. J. Biol. Chem. 274, 11549-11556.

Kuopio University Publications D. Medical Sciences D 434. Hassinen, Maija. Predictors and consequences of the metabolic syndrome: population-based studies in aging men and women. 2008. Acad. Diss. D 435. Saltevo, Juha. Low-grade inflammation and adiponectin in the metabolic syndrome. 2008. 109 p. Acad. Diss. D 436. Ervasti, Mari. Evaluation of Iron Status Using Methods Based on the Features of Red Blood Cells and Reticulocytes. 2008. 104 p. Acad. Diss. D 437. Muukka, Eija. Luomun tie päiväkotiin: luomuruokailun toteutettavuus ja ravitsemuksellinen merkitys päiväkotilapsille. 2008. 168 p. Acad. Diss. D 438. Sörensen, Lars. Work ability and health-related quality of life in middle-aged men: the role of physical activity and fitness. 2008. 83 p. Acad. Diss. D 439. Maaranen, Päivi. Dissociation in the finnish general population. 2008. 97 p. Acad. Diss. D 440. Hyvönen, Juha. Suomen psykiatrinen hoitojärjestelmä 1990-luvulla historian jatkumon näkökulmasta. 2008. 279 p. Acad. Diss. D 441. Mäkinen, Heidi. Disease activity and remission in rheumatoid arthritis: comparison of available disease activity measures and development of a novel disease sctivity indes: the mean overall index for rheumatoid arthritis (MOI-RA). 2008. 129 p. Acad. Diss. D 442. Kousa, Anne. The regional association of the hardness in well waters and the incidence of acute myocardial infarction in rural Finland. 2008. 92 p. Acad. Diss. D 443. Olkku, Anu. Glucocorticoid-induced changes in osteoblastic cells: cross-talk with wnt and glutamate signalling pathways. 2009. 118 p. Acad. Diss. D 444. Mattila, Riikka. Effectiveness of a multidisciplinary lifestyle intervention on hypertension, cardiovascular risk factors and musculoskeletal symptoms. 2009. 92 p. Acad. Diss. D 445. Hartmann-Petersen, Susanna. Hyaluronan and CD44 in epidermis with special reference to growth factors and malignant transformation. 2009. 103 p. Acad. Diss. D 446. Tolppanen, Anna-Maija. Genetic association of the tenomodulin gene (TNMD) with obesity- and inflammation-related phenotypes. 2009. 111 p. Acad. Diss. D 447. Lehto, Soili Marianne. Biological findings in major depressive disorder with special reference to the atypical features subtype. 2009. 115 p. Acad. Diss. D 448. Nieminen, Jyrki. Effect of functional loading on remodelling in canine, and normal and collagen type II transgenic murine bone. 2009. 107 p. Acad. Diss. D 449. Torpström, Jaana. Yliopistokoulutus ravitsemusasiantuntijuuden kehittäjänä. 2009. 164 p. Acad. Diss.