01.12.2012 Views

Matthias Kresse - KOPS - Universität Konstanz

Matthias Kresse - KOPS - Universität Konstanz

Matthias Kresse - KOPS - Universität Konstanz

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Mechanism of liver cell injury<br />

by the cytostatic drug melphalan<br />

DISSERTATION<br />

zur<br />

Erlangung des akademischen Grades eines<br />

Doktors der Naturwissenschaften (Dr. rer. nat.)<br />

des Fachbereiches für Biologie<br />

an der<br />

<strong>Universität</strong> <strong>Konstanz</strong><br />

vorgelegt<br />

von<br />

<strong>Matthias</strong> <strong>Kresse</strong><br />

Tag der mündlichen Prüfung: 11. Februar 2005<br />

1. Referent: Prof. Dr. A. Wendel<br />

2. Referent: Prof. Dr. A. Bürkle<br />

<strong>Konstanz</strong>, im Februar 2005


Acknowledgement<br />

This thesis was prepared between November 2000 and September 2004 at the Chair of<br />

Biochemical Pharmacology of the University of <strong>Konstanz</strong> under the instruction and<br />

supervision of Prof. Dr. A. Wendel.<br />

My special thanks go to my supervisor Prof. Wendel. He did not only provide excellent<br />

working facilities and a generous financing of the project, but also enabled this study by<br />

his patience, his encouragement, his criticism and stimulating ideas. Moreover he was<br />

always open to new ideas concerning projects or events and encouraged me to realise<br />

them by his support.<br />

This work benefited greatly from a number of excellent collaborations and discussions<br />

with other groups and within the DFG graduate trainee program “Biomedical Drug<br />

Research” directed by Prof. Dr. A. Wendel and Prof. Dr. K.P. Schäfer (ALTANA Pharma,<br />

<strong>Konstanz</strong>). This program enabled me to attend conferences, local and international<br />

training courses which provided a basis for fruitful discussions and insights in different<br />

fields of scientific and industrial interests of research. I am grateful for this support and<br />

the grant supplied by the graduate trainee program.<br />

I would like to extend my gratitude to Prof. Dr. R. Schmid for the excellent collaboration<br />

on the NFκB-work, to Dr. H.-M. Riehle for the help with histology and the interesting<br />

discussions, to Dr. N. Van Rooijen for providing Chlodronate Liposomes, to Dr. M.<br />

Menges for providing CD11b Microbeads, to Prof. Dr. K. Pfizenmaier for providing<br />

several mouse strains and to Prof. P. Scheurich for providing modified human fibroblasts.<br />

Very special thanks go to Prof. Dr. M. Schwarz and E. Zabinski for their great support in<br />

the tumor project and to Dr. N. Dikopoulos who has not only turned out to be a perfect<br />

partner in scientific joint-ventures, but also as a very good friend.<br />

Me and this work benefited also greatly from collaborations and discussions with a lot of<br />

members of ALTANA Pharma, namely Prof. Dr. K.-P. Schäfer, PD Dr. C. Schudt, Dr. R.<br />

Beume, Dr. M. Eltze, Dr. Th. Klein, Dr. Ch. Hesslinger, and K. Graf.<br />

Within the chair of Biochemical Pharmacology I am especially grateful to Dr. G. Künstle<br />

and Dr. R. Lucas for their supervision of the project and their help in technical questions,<br />

Dr. Th. Meergans for his advise and support, Dr. C. Hermann for her help concerning<br />

endotoxin and last but not least M. Weiller, Ch. Hart, Dr. M. Latta, Dr. C. Braun, Dr. K.<br />

Eichert and Dr. Markus Müller for their friendship, support and valuable discussions.<br />

Further thanks go to A. Hildebrandt and U. Gebert for their excellent technical<br />

assistance in protein analysis, ELISA-technique and animal experiments and to Gudrun<br />

Kugler for her never-ending patience and support in all kinds of daily problems.<br />

Finally, on a personal note, I would like to thank my family and my friends for their<br />

everlasting support that I’m very grateful for.<br />

<strong>Konstanz</strong>, September 2004 <strong>Matthias</strong> <strong>Kresse</strong>


Table of contents<br />

1 Introduction _____________________________________________________________10<br />

1.1 From battlefield to bedside – History of chemotherapy ____________________________10<br />

1.2 Alkylating agents ___________________________________________________________11<br />

1.3 Mechanism of action and cytotoxicity __________________________________________13<br />

1.3.1 Cyclophosphamide _______________________________________________________________ 15<br />

1.3.2 Melphalan ______________________________________________________________________ 17<br />

1.4 Chemical and pharmacological aspects of melphalan______________________________18<br />

1.5 Nitrogen mustards - Clinical aspects of tumor-therapy ____________________________19<br />

1.5.1 Systemic application of melphalan __________________________________________________ 20<br />

1.5.2 Isolated limb and liver perfusion____________________________________________________ 21<br />

1.6 TNF – heaven and hell in cancer therapy _______________________________________26<br />

1.7 A whole family of cytotoxic cytokines – the TNF superfamily _______________________27<br />

1.7.1 CD95, Fas and APO-1 ____________________________________________________________ 29<br />

1.7.2 TRAIL _________________________________________________________________________ 30<br />

1.8 Combinatorial therapy_______________________________________________________31<br />

1.8.1 TNF – the good, the bad or the ugly in ILP ___________________________________________ 32<br />

1.8.2 Investigation of hyperthermia in ILP ________________________________________________ 32<br />

1.8.3 Pre-clinical results of melphalan in cancer therapy_____________________________________ 34<br />

1.9 Apoptosis __________________________________________________________________35<br />

1.9.1 Signaling of TNF receptors in apoptosis and anti-apoptosis______________________________ 36<br />

1.9.2 The role of TNF receptor 1 ________________________________________________________ 37<br />

1.9.3 The role of TNF receptor 2 ________________________________________________________ 38<br />

1.9.4 Anti-apoptotic signaling ___________________________________________________________ 39<br />

1.9.5 Modulation of apoptosis by changing cellular redox and energy state _____________________ 40<br />

1.9.6 Differences in apoptosis of primary versus transformed cells ____________________________ 42<br />

1.9.7 Apoptosis in cancer_______________________________________________________________ 44<br />

2 Objectives of the thesis _____________________________________________________46<br />

3 Materials and methods _____________________________________________________47<br />

Materials and animals _____________________________________________________________47<br />

3.1 Chemicals _________________________________________________________________47<br />

3.2 Antibodies and recombinant proteins __________________________________________47<br />

3.3 Cell culture materials________________________________________________________48


Table of contents<br />

3.4 Animals ___________________________________________________________________48<br />

Methods _________________________________________________________________________49<br />

3.5 Cell culture ________________________________________________________________49<br />

3.5.1 Cell lines________________________________________________________________________ 49<br />

3.5.2 Isolation and culture of primary cells ________________________________________________ 50<br />

3.6 Isolated liver perfusion ______________________________________________________51<br />

3.6.1 Technical procedure of liver preperation and perfusion_________________________________ 51<br />

3.6.2 Treatment schedules______________________________________________________________ 53<br />

3.6.3 Sampling of material _____________________________________________________________ 54<br />

3.7 Animal experiments _________________________________________________________54<br />

3.7.1 Treatment schedules______________________________________________________________ 54<br />

3.7.2 Sampling of material _____________________________________________________________ 55<br />

3.8 Tumor model_______________________________________________________________56<br />

3.8.1 Tumor induction _________________________________________________________________ 56<br />

3.8.2 Isolated liver perfusion____________________________________________________________ 56<br />

3.8.3 Treatment schedules______________________________________________________________ 56<br />

3.9 Cytokine determination ______________________________________________________57<br />

3.10 Light microscopy ___________________________________________________________58<br />

3.10.1 HE-staining _____________________________________________________________________ 58<br />

3.10.2 Tunnel-staining __________________________________________________________________ 58<br />

3.11 Measurement of enzyme activities _____________________________________________58<br />

3.11.1 Liver enzyme activities in plasma or samples _________________________________________ 58<br />

3.11.2 Lactate dehydrogenase activity _____________________________________________________ 59<br />

3.11.3 Caspase 3,7-like activity ___________________________________________________________ 59<br />

3.11.4 TNF-α-converting enzyme (TACE)-activity___________________________________________ 60<br />

3.12 Determination of ATP _______________________________________________________60<br />

3.13 Determination of glutathione__________________________________________________61<br />

3.14 Cytokine determination ______________________________________________________61<br />

3.15 Cell-based ELISA___________________________________________________________61<br />

3.16 Western blot analysis ________________________________________________________62<br />

3.17 Statistics___________________________________________________________________63<br />

4 Results__________________________________________________________________64<br />

4.1 Characterisation of melphalan-induced cell death in primary murine liver cells _______64<br />

4.2 The role of death receptors in melphalan-induced apoptosis________________________67


Table of contents<br />

4.3 TNF mediates melphalan-induced toxicity ______________________________________69<br />

4.4 Kupffer cells are the source of TNF mediating melphalan-induced toxicity ___________70<br />

4.5 Kupffer cells lead to enhanced TNF secretion after LPS but not after melphalan<br />

stimulation_________________________________________________________________72<br />

4.6 Melphalan inhibits recombinant TACE activity in vitro ___________________________76<br />

4.7 Melphalan is able to block TNF-dependent liver injury in vivo _____________________77<br />

4.7.1 GalN/TNF and GalN/LPS model____________________________________________________ 77<br />

4.7.2 GalN/SEB ______________________________________________________________________ 80<br />

4.7.3 The Con A model. ________________________________________________________________ 81<br />

4.7.4 The CD95 model _________________________________________________________________ 83<br />

4.7.5 Melphalan and melphalan/TNF model _______________________________________________ 84<br />

4.8 Cell-cell contact is necessary between Kupffer cells and hepatocytes _________________84<br />

4.9 Role of TNF receptors on Kupffer cells _________________________________________85<br />

4.10 Melphalan-induced toxicity on Kupffer cells_____________________________________87<br />

4.11 Melphalan-induced hepatotoxicity in the presence and absence of Kupffer cells _______88<br />

4.12 Melphalan induces TNF-mediated hepatotoxicity in situ ___________________________90<br />

4.13 Modulation of intracellular signaling by various inhibitors_________________________94<br />

4.13.1 Role of NFκB____________________________________________________________________ 94<br />

4.13.2 Role of p38 MAPK _______________________________________________________________ 96<br />

4.13.3 Role of JNK _____________________________________________________________________ 97<br />

4.13.4 Role of PARP____________________________________________________________________ 98<br />

4.14 Influence of melphalan on cellular redox and energy state and vice versa ____________100<br />

4.14.1 Melphalan and cellular redox state _________________________________________________ 100<br />

4.14.2 Melphalan and cellular energy status _______________________________________________ 101<br />

4.15 Effect of melphalan on DEN-induced liver tumors in isolated liver perfusion of tumorbearing<br />

mice ______________________________________________________________106<br />

4.15.1 DEN induced massive tumor growth in murine livers following intraperitoneal injection ____ 106<br />

4.15.2 Isolated liver perfusion of tumor-bearing livers with melphalan/TNF ____________________ 108<br />

4.15.3 Activation of caspases in normal and neoplastic tissue _________________________________ 109<br />

4.15.4 Histological examination of perfused livers derived from tumor-bearing mice _____________ 109<br />

5 Discussion______________________________________________________________112<br />

5.1 The role of TNF and TNF receptors in melphalan-mediated hepatotoxicity __________113<br />

5.2 Poseidon’s trident - mechanism of melphalan-induced cytotoxicity _________________115<br />

5.3 Melphalan-induced upregulation of membrane TNF _____________________________118


Table of contents<br />

5.4 Clinical relevance __________________________________________________________119<br />

5.5 Differential role of TNF receptors in the presence and absence of Kupffer cells_______120<br />

5.6 Intracellular signaling of TNF receptors following melphalan incubation____________122<br />

5.6.1 Nuclear factor κ B_______________________________________________________________ 123<br />

5.6.2 p38 mitogen activated protein kinase and c-Jun N-terminal kinase ______________________ 124<br />

5.7 Modification of melphalan-toxicity by altered cellular redox or energy state _________126<br />

5.7.1 Glutathione ____________________________________________________________________ 126<br />

5.7.2 ATP __________________________________________________________________________ 127<br />

5.8 Influence of ATP depletion on melphalan/TNF-mediated organ toxicity in tumor-bearing<br />

mice _____________________________________________________________________129<br />

6 Summary_______________________________________________________________131<br />

7 Zusammenfassung _______________________________________________________133<br />

8 References______________________________________________________________135


αCD95 agonistic anti-CD95-antibody<br />

ActD actinomycin-D<br />

ALT alanin-aminotransferase<br />

AP-1 activator protein 1<br />

Apaf-1 apoptosis activating-factor 1<br />

ATP adenosintriphosphate<br />

bp basepair<br />

BIR baculoviral inhibitory repeat<br />

BSA bovine serum albumin<br />

BSO buthionine-sulfoximine<br />

CARD caspase recruitment domain<br />

CD cluster of differentiation<br />

CD95 Fas, Apo-1<br />

CD95L CD95-ligand<br />

CHX cycloheximide<br />

Con A Concanavalin A<br />

CR complete response<br />

CTL cytotoxic T lymphocyte<br />

Cyt c cytochrome c<br />

DD death domain<br />

DED death effector domain<br />

DEM diethylmalate<br />

DEN diethlnitrosamine<br />

DEVDafc N-acetyl-asp-glu-val-asp-7-amino-4trifluoromethyl<br />

coumarin<br />

DISC death-inducing signaling complex<br />

EDTA ethylendiamintetraacetic acid<br />

ELISA enzyme-linked immunosorbent assay<br />

FADD Fas-associated protein with death<br />

domain<br />

FAS FS-7-associated surface antigen<br />

FCS foetal calf serum<br />

FLICE FADD-like ICE<br />

GalN D-galactosamine<br />

GPT glutamate-pyruvate-transaminase<br />

GSH glutathione (reduced state)<br />

GSSG glutathione disulphide<br />

GSx glutathione (total)<br />

HAI hepatic artery infusion<br />

HC hepatocyte<br />

HCC hepatocellular carcinoma<br />

HEPES N-2hydroxyethylpiperazi-N’-2ethansulfonic<br />

acid<br />

HSA human serum albumin<br />

HSP heat shock protein<br />

hu human<br />

i.p. intraperitoneally<br />

Abbreviations<br />

i.v. intravenously<br />

ICE interleucin-1ß-converting enzyme<br />

IFNγ interferon gamma<br />

IHP isolated hepatic perfusion<br />

IL interleukin<br />

ILP isolated limb perfusion<br />

IPML isolated perfused mouse liver<br />

JNK JunNH2-terminale kinase<br />

LAL Limulus amoebocyte lysate<br />

LDH lactate dehydrogenase<br />

LPS lipopolysaccharide<br />

mAb monoclonal antibody<br />

MOF multi-organe failure<br />

mu murine<br />

MΦ macrophage<br />

NCS new-born calf serum<br />

NFκB nuclear factor kappa B<br />

NIK NFκB-inducing kinase<br />

NPCs non-parenchymal liver cells<br />

ORR overall response rate<br />

p55 Tumor necrosis factor receptor 1<br />

p75 Tumor necrosis factor receptor 2<br />

L-PAM L-phenylalanine mustard<br />

PAGE polyacrylamide gel electrophoresis<br />

PBS phosphate-buffered saline<br />

PCD programmed cell death<br />

PR partial response<br />

RIP Receptor-interacting protein<br />

RAIDD RIP-associated ICE-homologous protein<br />

with death domain<br />

S.E.M. standard error of means<br />

SD standard derivation<br />

SDS sodiumdodeclysulfate<br />

SEC sinusoidal endothelial cells<br />

TLR TOLL-like receptor<br />

TNF tumor necrosis factor<br />

TNFR1 tumor necrosis factor receptor 1<br />

TNFR2 tumor necrosis factor receptor 2<br />

TRADD TNF receptor-associated protein with<br />

death domain<br />

TRAF2 TNF receptor-associated factor 2<br />

TRAIL TNF-related apoptosis inducing ligand<br />

Tris tris-(hydroxymethyl-)aminomethan<br />

vs. versus<br />

w/o without<br />

zVAD-fmk benzyloyxcarbonyl-val-ala-asp-(OMe)fluoromethylketone


1 Introduction<br />

Introduction<br />

1.1 From battlefield to bedside – History of chemotherapy<br />

During the First World War a new kind of warfare agent was developed by German<br />

chemists, the so-called sulphur mustards, also called LOST in remembrance of their<br />

discoverer Lommel and Steinkopf. The first well-documented military uses took place on<br />

July 12 th 1917 in Ieper (Flanders, Belgium) and Abessinia (Italy) 1935-36 (1), but there is<br />

also evidence that a kind of sulphur mustards had been used in the 4 th century by<br />

Chinese commanders on the battlefield blowing sulphur mustards against their enemies<br />

by the use of fans. In Belgium in the 20 th century at least 70 soldiers were killed, 2000<br />

very heavily poisoned and the number of chronic injuries has never been established.<br />

Sulphur mustards were observed to be taken up very easily through the skin<br />

because of their strong lipophilic behaviour. Subsequently these substances cause<br />

severe lung oedema, irritations of eyes and skin, followed by severe leucopoenia, bone<br />

marrow aplasia, dissolution of lymphoid tissues and severe ulceration in the<br />

gastrointestinal tract. Long term consequences in liver and kidney have still been<br />

observed for decades later. These findings, which suggest that the toxic effect of these<br />

substances targets rapidly dividing cells, raised the idea that there could be a<br />

pharmacological implementation in diseases characterised by uncontrolled cell<br />

proliferation, such as cancer.<br />

The first clinical trial in cancer therapy was published in 1931. A Sulphur mustard<br />

solution was applied topically onto or injected directly into human tumors (2). The<br />

procedure was subsequently abandoned due to the high toxicity in the patients, but the<br />

attempts to cure cancer with substances of this type continued. It was however for more<br />

than ten years later in 1946 that Gilman and Philips published their initial results of<br />

successful clinical trials with this new kind of substances (3). This work is referred to as<br />

the beginning of modern cancer therapy.<br />

Since those early works in the field of chemotherapeutic agents, hundreds of<br />

different substances have been developed and used clinically. Most of them target the<br />

DNA by direct chemical reaction (alkylating agents, platinum compounds), by synthesis<br />

inhibition (topoisomerase inhibitors, anti-metabolites) or by interaction in the cellular<br />

machinery during cell division (taxanes, vinca-alkaloids). Consequently, all anti-cancer<br />

drugs display a high activity against proliferating cells or tissues with a large proportion<br />

of dividing cells. Unfortunately, most tumors have proportions of dividing cells very<br />

similar to what is found in regenerating or strongly proliferating healthy tissues, like the<br />

- 10 -


Introduction<br />

bone marrow, hair follicles, germ cells, the gastrointestinal mucosa and the liver,<br />

contributing to a wide range of side-effects mostly observed with the use of these<br />

compounds. In recent years, this topic attracted much attention and many investigations<br />

have been done to develop clinical strategies, in order to prevent general toxicity,<br />

without interfering with the toxic effects on the tumor itself. In addition to chemotherapy,<br />

surgery and radiotherapy are corner-stones in the treatment of malignancies. Each of<br />

these modalities is associated with specific risks and benefits, according to the individual<br />

prognosis of the patient and therefore modern cancer therapy often combines different<br />

treatment strategies with the aim to concentrate the cytotoxic properties at the tumor<br />

site, while diluting the unwanted. In contrast to surgery and radiotherapy, which can be<br />

considered as a local or regional regimen, chemotherapy is mostly given systemically,<br />

offering a treatment strategy that potentially reaches metastatic cells, even if their<br />

existence is far away or unknown. On the other hand, systemic application is often<br />

accompanied with more extensive and systemic toxicity, leading to the severe sideeffects<br />

mentioned above. To solve this conflict, several strategies have been developed,<br />

which combine surgery and chemotherapy in a regional treatment regimen. However,<br />

despite of all advantages these new techniques and strategies offer, no strategy has<br />

been developed to distinguish between the tumor cell and the neighbouring healthy cell.<br />

This thesis focuses on the molecular basis of cell death induction in primary and<br />

transformed liver cells by melphalan, an alkylating agent derived from nitrogen mustard.<br />

The interest in this topic resulted from the expanded knowledge in tumor biology and cell<br />

death during the last decades, especially in the field of induction and modulation of<br />

apoptosis in normal versus transformed tissue.<br />

1.2 Alkylating agents<br />

Alkylating agents are compounds that react with electron-rich atoms in biologic<br />

molecules, to form covalent bonds. Traditionally these compounds have been classified<br />

into two types, on the one hand compounds that react directly with biological molecules<br />

and follow a so-called SN1-mechanism and on the other hand compounds that form a<br />

reactive intermediate, which then reacts with the biologic molecule (SN2-mechanism).<br />

A large number of chemical compounds have been found to alkylate biological<br />

molecules under physiological conditions, but the most frequently used among them are<br />

derivatives of nitrogen mustards. In the last century hundreds of different nitrogen<br />

mustards have been synthesised and tested, but only five of them are commonly used in<br />

- 11 -


Introduction<br />

cancer therapy today. These are mechlorethamine, cyclophosphamide, ifosfamide,<br />

chlorambucil and melphalan (L-PAM), as illustrated in figure 1.<br />

- 12 -


Introduction<br />

1.3 Mechanism of action and cytotoxicity<br />

Nitrogen mustards exert their cytotoxic action through covalent interaction with<br />

intracellular nucleophils, especially DNA, as a result of spontaneous formation of a<br />

reactive cyclic intermediate, the aziridinium ion (figure 2). Bifunctional agents are much<br />

more effective anti-cancer drugs than monofunctional ones, because these compounds<br />

are able to cross-link a DNA strand within a double helix (intrastrand) (figure 2), between<br />

two strands (interstrand) or between DNA and proteins.<br />

Cross-linking of DNA is probably the most important factor for the cytotoxic effect,<br />

resulting in inhibitory effects on DNA replication and transcription, which subsequently<br />

leads to cell death. It has been shown that nitrogen mustards exert their antitumor<br />

activity only when the two-armed mustard is intact (4). The reactivity of the nucleophilic<br />

sites in the DNA relates to electronic and steric properties, as well as to the involvement<br />

of hydrogen bonds within the double helix.<br />

- 13 -


Introduction<br />

Detailed studies have revealed that the N-7 position of guanylic acid is the most<br />

susceptible site for alkylation (5). Brookes and Lawley also suggested that the nitrogen<br />

mustard cross-link was between the N-7 guanine atoms in base-paired G-C sequences<br />

in the DNA (6). Bauer and Povirik could demonstrate that mustards form two basestaggered<br />

interstrand cross-links between the 5’ G and the G opposite C in the 5’ G-G-C<br />

sequence in double stranded DNA of CHO cells in vitro (7). In contrast to melphalan,<br />

mechlorethamine and phosphoramide were also able to form intrastrand crosslinks<br />

between two contiguous Gs in the G-G-C sequence in double stranded DNA. It was<br />

shown that melphalan is only able to cause intrastrand cross-links in single stranded<br />

DNA with a very slow kinetics of second arm alkylation and therefore cross-links<br />

between two DNA strands can only be formed when monoadducted DNA becomes<br />

transiently single stranded during transcription or replication (7).<br />

Another interesting finding is that patients treated with nitrogen mustards had an<br />

increased risk for secondary myeloma (8). Although the exact mechanism of this finding<br />

remains largely unknown, it was found that aromatic nitrogen mustards like melphalan<br />

can also lead to A-T→T-A transversions, apparently resulting from adenine N-3<br />

alkylations (9, 10).<br />

But alterations of the DNA are not the only effect nitrogen mustards cause.<br />

Investigations on the mechanism of primary lesions caused by nitrogen mustards, lead<br />

to the so called “Papirmeister-hypothesis” (11). In response to DNA damage, the cell<br />

cycle is arrested and a class of enzymes called poly-adenosinediphosphate-ribosepolymerase<br />

(PARP) (other names in literature are poly(ADP-ribose) synthetase (PARS)<br />

or poly(ADP-ribose) transferase (pADPRT)) (12), which is abundantly present in the<br />

nucleus (13, 14), is activated. The obligatory triggers of PARP activation are nicks and<br />

breaks in the DNA strand, which can be induced by a broad variety of environmental<br />

stimuli including alkylating agents (12). On the average, one molecule of this enzyme is<br />

present per 1000 bp of DNA in the nucleus. Upon activation, the enzyme uses<br />

nicotinamide-adenine-dinucleotide (NAD) as a cellular substrate to build up<br />

homopolymers of adenine diphosphate ribose units. In case of massive DNA alkylation,<br />

this leads to a massive decrease of cellular NAD-pools which, in turn leads to inhibition<br />

of glycolysis and energy metabolism, to activation of monophosphate shunts, to the<br />

breakdown of the cellular redox status, the release of proteases and finally to cell death<br />

(15, 16).<br />

Evidence for this phenomenon also comes from a totally independent field of<br />

scientific interest. In 1999, Burkart and co-workers published their data on an<br />

experimental model for type 1 diabetes due to the loss of pancreatic beta-cells. The<br />

- 14 -


Introduction<br />

authors demonstrated that PARP -/- mice were completely protected from streptozocininduced<br />

diabetes (17). Streptozocin (also: Streptozotoxin) is able to induce double<br />

strand breaks in islet cell DNA upon uptake via glucose-transporters type 2 GLUT-2 (16,<br />

17). PARP was shown to be activated and cellular NAD-level decreased to a very low<br />

level, resulting in abrogation of sufficient energy generation and finally beta-cell death<br />

(for illustration see also (18). In PARP -/- mice this effect was totally abolished and mice<br />

were completely protected from Streptozocin-induced hyperglycaemia.<br />

Nitrogen mustards and alkylating agents in general can also react spontaneously<br />

with glutathione, but this aspect will be discussed later more detailed. In the following<br />

section the mechanism of two antitumor compounds, cyclophosphamide and melphalan,<br />

will be discussed in detail, because of their greater clinical relevance during the last<br />

decades.<br />

1.3.1 Cyclophosphamide<br />

Cyclophosphamide is not toxic itself, but it undergoes activation in the liver which<br />

finally leads to a toxic product. The initial activation reaction is carried out by the<br />

microsomal oxidation of cytochrome P450 (CYP2B6) producing 4hydroxycyclophosphamide,<br />

which is predominantly present in that form at physiologic<br />

pH, but which is also in equilibrium with aldophosphamide. This equilibrium mixture is<br />

able to diffuse from the hepatocyte into the plasma and is distributed throughout the<br />

body. Since 4-hydroxy-cyclophosphamide is relatively nonpolar, it is able to enter target<br />

cells in the periphery readily by diffusion. Aldophosphamide instead decomposes<br />

spontaneously to produce phosphamide mustard (for illustration see also figure 3),<br />

which in turn is able to react spontaneously with DNA or proteins as described by other<br />

sulphur- or nitrogen-mustards.<br />

- 15 -


Introduction<br />

Phosphamide-mustard is also produced outside the cell, but this compound is<br />

very polar and enters cells poorly. The intracellular toxification of cyclophosphamide was<br />

of great advantage for clinical investigations. It can be administrated in much higher<br />

doses (up to 170 mg/kg as continuously infusion over 4 days or 110 mg/kg over 90<br />

minutes), as compared to melphalan (0.6 mg/kg intravenously or 0.25 mg/kg orally) for<br />

example. But unlike melphalan, patients treated with cyclophosphamide displayed much<br />

less side effects.<br />

- 16 -


1.3.2 Melphalan<br />

Introduction<br />

Melphalan was first synthesised by two British scientists, i.e. Bergel and Stock, in<br />

the early 1950’s. The two scientists believed that, based on previous results with serinealanine-<br />

and phenylalanine derivatives, certain natural amino acids or peptides might<br />

show antitumor activity, possibly by interfering with the nucleic acid or protein<br />

metabolism of malignant cells (19-21). The first publication focused on derivatives of<br />

phenylalanine, in which three forms of bis(2-chloroethyl)aminophenylalanine were<br />

described, the D-, L- and the racemic D-L-form (table 1). Subsequent testing of these<br />

substances in a Walker rat carcinoma model revealed that only the L-form had a high<br />

antitumor activity, while the D- and the racemic DL-form was much less active. The lead<br />

structure of p-L- bis(2-chloroethyl)aminophenylalanine was later named “melphalan”,<br />

derived from mustard –L-phenylalanine. According to this nomenclature, the D-form was<br />

named medphalan and the racemic DL-form merphalan.<br />

- 17 -


Introduction<br />

In parallel to Bergel and Stock, some Russian scientists Larionov and colleagues<br />

worked on a different compound which was named “sarcolysine” (table 1), because of its<br />

remarkable effect on soft tissue sarcomas (22). Later on it was found that sarcolysine<br />

and merphalan were exactly the same substances and therefore had been often referred<br />

to as merphalan. The racemic compound was only used clinically in Eastern Europe for<br />

a period of time, but never reached the same international interest, as compared to the<br />

L-enantiomer. The reasons for this remain unclear. But it could be related to early<br />

observations of Bergel and Stock, who demonstrated a better cellular uptake of the Lform<br />

in vivo.<br />

In 1965 Schmidt and colleagues published a series of in vivo toxicity experiments<br />

conducted at the US Cancer Chemotherapy National Service Center (23-25). Those<br />

studies focused on the toxicity and anti-tumor activity of 20 reference substances, 39<br />

bis(2-chloroethyl)amino-derivatives, 20 aziridines and 35 methanesulfonates in mice and<br />

rats. For L-forms of phenylalanine mustards a clear tendency for increased toxicity was<br />

found compared to D-forms. Therefore, not surprisingly, the L-enantiomers were clearly<br />

favoured.<br />

1.4 Chemical and pharmacological aspects of melphalan<br />

Melphalan is a bifunctional alkylating agent, chemically known as 4-[bis(2chloroethyl)amino]-L-phenylalanine,<br />

with a molecular formula of C13H18Cl2N2O2 and a<br />

molecular weight of 305.20 g/mol. Melphalan is a white odourless powder that is almost<br />

insoluble in water (pKa of ~2.5), but which is subject to rapid hydrolysis in water and<br />

plasma. It is soluble in ethanol, propylene glycol, alkaline solution, 2%<br />

carboxymethylcellulose and diluted mineral acid. The half-life of melphalan in a<br />

physiological buffer at 37°C is approximately 1.5 hours, but the stability in solution is<br />

influenced by pH, temperature and protein content (plasma proteins) of the solvent and<br />

of the concentration of the drug. In general, more dilute solutions of melphalan degrade<br />

slightly quicker (26). For NaCl-solutions it was found that melphalan is 30% more stable<br />

in 150 mM NaCl solution (normal saline), as compared to Dulbecco’s phosphatebuffered<br />

saline. Stored in normal saline at room temperature the drug lost 5% of its<br />

activity (t0.95) in 1.5 hours, at 4°C the t0.95 was, however, 20 hours (26). In addition to that<br />

Flora, Smith and Cradock reported that hydrolysis of melphalan is inhibited by increasing<br />

chloride ion concentrations (27). In an aqueous solution of 28°C an 80% decomposition<br />

was observed after 24 hours.<br />

- 18 -


Introduction<br />

The pharmacokinetics of intravenously and orally administered melphalan has<br />

been extensively studied in humans. For all other species hardly any experimental data<br />

are available. Following injection drug plasma concentrations declined rapidly in a biexponential<br />

manner. With the distribution phase and terminal elimination phase half-lives<br />

of approximately 75 minutes were calculated. Melphalan is eliminated from plasma<br />

primarily by chemical hydrolysis to mono- and dihydroxy derivatives. Other metabolites<br />

have not been observed in humans. Melphalan is able to pass the blood-brain-barrier,<br />

but the penetration into the cerebrospinal fluid is low. Interestingly, there is evidence that<br />

nitrogen mustards enter the cell via an active cellular uptake since Goldenberg et al.<br />

have been demonstrated that there is evidence for a transport carrier of nitrogen<br />

mustard in nitrogen mustard-sensitive and -resistant L5178Y lymphoblasts. This<br />

assumption was based on the finding that the time course for the uptake of labelled<br />

nitrogen mustard was much more rapid in sensitive lymphoblasts and with a higher<br />

extends compared to resistant lymphoblasts. Additionally cellular uptake could be<br />

inhibited by 2,4-dinitrophenol (DNP, 1 mM) and oubain (0.5 mM) (28, 29). The extent of<br />

melphalan binding to plasma protein is high, ranging from 60% to 90%. After injection of<br />

radiolabelled melphalan, radioactivity is excreted both in urine and faeces (30).<br />

1.5 Nitrogen mustards - Clinical aspects of tumor-therapy<br />

Since 1946, when the first clinical trial with sulphur mustards was presented by<br />

Gilman and Philips a new kind of therapy for patients with cancer concerning various<br />

organs has been born. At the beginning of the 1960s melphalan was the first drug that<br />

improved the therapy of multiple myeloma. The median of survival was prolonged from<br />

several months to 3 years. In the following four decades several new drugs were tested<br />

including cisplatin, mitomycin c and 5-fluorouracil. But no other drug brought better<br />

results than melphalan (31). From 1950 till today numerous articles and clinical studies<br />

have been reported, demonstrating enhanced success with different application forms,<br />

clinical protocols and combinatorial therapies of nitrogen mustards, cytokines and<br />

hyperthermia. Table 2 gives an overview of the milestones that will be discussed in the<br />

next sections.<br />

- 19 -


Introduction<br />

1.5.1 Systemic application of melphalan<br />

Therapy with melphalan and prednisone in intermittent course has represented the<br />

“golden standard” for newly diagnosed symptomatic multiple myeloma for many years<br />

(32). Approximately 50% of treated patients responded, but the frequency of complete<br />

remission was rather low (10%) (33). In addition the median remission time was only two<br />

years and the median survival three years (31). A disease-free long term survival could<br />

only be accomplished by allogeneic bone marrow transplantation.<br />

- 20 -


Introduction<br />

Comparative studies have been started to investigate higher response rates by<br />

polychemotherapy, but none of the treatment modalities differed in the survival<br />

parameters (31). The so-called VAD regimen (infusion of vincristin and doxorubicin and<br />

oral dexamethasone) turned out to be an effective treatment schedule for melphalan<br />

non-responders (34), but the overall survival was not enhanced compared to melphalanprednisone<br />

treatment (35). Therefore, the transience of enhanced remission by<br />

polychemotherapy and alternative VAD regimens led to an investigation of dose<br />

intensification. In 1983 McElwain reported results using a melphalan dose of 140 mg/m²<br />

for the treatment of patients with plasma cell leukaemia or multiple myeloma (36). The<br />

first promising results led to further investigations with even more dose intensification up<br />

to 200 mg/m² (37). But despite of higher response rates that could be reached by high<br />

dose drug treatment, the overall survival rates did not change dramatically. More than<br />

14% of patients died very early due to the high substance-related toxicity (38) and the<br />

morbidity was also very high (39).<br />

In view of the high systemic toxicity resulting from orally or intravenous application<br />

of cytostatic drugs, a new strategy for treatment of melanoma patients was under<br />

development since the 1950’s, using a local regimen in form of an isolated limb<br />

perfusion (ILP). Cancer treatment with ILP enabled high response rates, enhanced<br />

survival rates and low systemic toxicity due to the local application of nitrogen mustards.<br />

In the following section this idea of cancer treatment will be reviewed in more detail.<br />

1.5.2 Isolated limb and liver perfusion<br />

The most common application of isolated perfusion has been isolated limb<br />

perfusion for intransit-melanoma and as a limb-salvage technique for non-resectable or<br />

locally recurrent extremity sarcoma (40, 41). The perfusion circuit allows delivering<br />

hyperthermia, biological agents and/or chemotherapy via a closed recirculation usually<br />

for one or two hours (42). At the end of the treatment, the perfused organ is flushed with<br />

saline or colloid solution to remove any remaining traces of the therapeutics, before the<br />

vascular continuity is restored. This technical procedure is more or less common to all<br />

settings of isolated perfused organs. Especially for isolated liver perfusion, two different<br />

approaches have been published in literature, the isolated hepatic perfusion (IHP) and<br />

the hepatic artery infusion (HAI).<br />

Starting at the late 1950’s Ausman (43) and Healey (44) studied a new technique<br />

for an isolated hepatic perfusion in animals. The initial clinical study performed by<br />

- 21 -


Introduction<br />

Ausman in 1961 was limited to five patients and the treatment was restricted to one hour<br />

(43). But against all hopes, high morbidity and mortality due to this technically<br />

demanding procedure was observed and only a weak acceptance of this technique was<br />

gained. Today their work can nevertheless be considered as a milestone in clinical<br />

cancer therapy.<br />

The technique was intended, however, to improve the efficacy of high-dose<br />

chemotherapy within the tumor site, whilst keeping systemic drug levels low to reduce<br />

systemic toxicity. Subsequently a number of articles were published, reporting different<br />

strategies for isolated hepatic perfusion (IHP), concerning the mode of perfusion<br />

(transarterial vs. intravenously). The idea of hepatic artery infusion (HAI) resulted from<br />

the observation that liver metastases with a diameter smaller than 5 mm are nourished<br />

mainly by the hepatic artery (45). In addition to that some drugs are more effectively<br />

extracted by the liver during their first pass through its arterial circulation and many<br />

drugs have a steep dose-response curve, so that large dose given regionally will lead to<br />

a better response (46). IHP has several theoretical advantages over HAI. First, the<br />

concentration of the perfused drug can be higher (only the organ toxicity is limiting).<br />

Second, the drug must not necessarily have a high hepatic extraction rate on the first<br />

pass. Third, conditions acting synergistically with chemotherapy might be added, like<br />

hypoxia or hyperthermia. Last but not least additional biotherapy using TNF and IFNγ<br />

might be more efficient and safe (46-49). Studies performed in rats by Marinelli et al.<br />

favour this rationale by the finding that administration of mustards in isolated rat liver<br />

produced higher tumor and tissue concentrations as compared to HAI (50, 51).<br />

In 1983 and 1986 Skibba and co-workers published their results using a four hour<br />

IHP under hyperthermic conditions without application of alkylating agents (52, 53).<br />

Unfortunately, within those publications no classical response criteria were mentioned.<br />

In 1994 Hafström and colleagues published data from their two clinical trials. In the first<br />

paper, 29 patients suffering from liver metastases originating from various primary<br />

lesions (breast cancer, colorectal cancer, midgut carcinoids and miscellaneous<br />

primaries) were treated with cisplatin and melphalan under hyperthermic conditions. The<br />

authors stated that four patients died within 30 days of multiple organ failure and partial<br />

tumor regression was only observed in 20% of the patients. All surviving participants<br />

developed reversal liver and renal toxicity (54). In their second paper a clinical study<br />

including 32 patients with cancer confined to the liver was referred. Patients were<br />

treated with 5-fluorouracil in a local regimen. Patients undergoing IHP displayed a<br />

median survival of 17 months, while two patients died due to toxicity from the wrong<br />

dose of 5-FU and the wrong route of administration (55). In 1996, Marinelli and coworkers<br />

reported an overall response rate of 25% and a median survival of 17 months<br />

- 22 -


Introduction<br />

using mitomycin c (56). In the same year Leff et al. reported a study of 20 patients with<br />

metastatic colon carcinoma treated with 180 mg/m² melphalan intravenously and<br />

subsequent autologous bone marrow transplantation. This study reached an overall<br />

response rate of 45% and a median survival of 198 days (57). Two years later, Aigner<br />

and colleagues reported first results of IHP using 300 mg and 450 mg 5-fluorouracil. In<br />

both patient groups several metastases had disappeared and the remaining tumor tissue<br />

showed extended necrotic areas.<br />

In 1998, De Vries and colleagues published a paper stating results of pre-clinical<br />

use of melphalan combined with exogenous TNF in isolated livers of guinea pigs and an<br />

additional phase I clinical study including nine patients. A partial response rate was<br />

reported for five patients, one patient developed stable disease and three patients died<br />

due to operation. The design of the clinical protocol included the cannulation of the<br />

artery and the portal vein for IHP with melphalan and TNF being used simultaneously for<br />

cancer treatment (58). Thought this trial was very successful on the one hand showing<br />

an overall response rate of nearly 100 %, new problems became obvious on the other<br />

hand. The combination of melphalan and recombinant TNF for therapy was often<br />

descibed to be associated with higher liver toxicity, fatal coagulative disturbances and<br />

severe systemic toxicity based on operative leakage of the perfusate.<br />

The problem of capillary leak during isolated hepatic perfusion was also discussed<br />

in a paper by Alexander et al. published in 1998. The authors presented a clinical study<br />

evaluating TNF as a potential mediator of capillary leak occurring during IHP. In that<br />

study 27 patients underwent a 60 min hyperthermic IHP with 1,5 mg melphalan, either<br />

combined with (n=7) or without (n=20) 1.0 mg of recombinant TNF. The study revealed<br />

that no significance could be calculated between both groups concerning leakage of<br />

radiolabelled albumin. Moreover TNF did not seem to be responsible for melphalan<br />

uptake in tumor cells. Thus leading the authors to the conclusion that the cytokine TNF<br />

has to be critically evaluated for clinical use in cancer treatment (59).<br />

More recently, being inspired by the promising report published for combinatorial<br />

melphalan and TNF in the isolated limb perfusion by Lienard et al. in 1992 (60), several<br />

institutions have been investigated combinatorial TNF/nitrogen mustard-based therapy.<br />

In their initial reports the authors observed a 90% complete response rate and an overall<br />

response rate of 100% in isolated limb perfusion using a combinatorial therapy with<br />

TNF, IFNγ and melphalan for patients with intransit melanoma or non-resectable<br />

extremity sarcoma (60, 61). This was considerably better than what had been reported<br />

previously by Alexander et al., using melphalan alone (62). The effects of these initial<br />

reports could be largely confirmed by others (reviewed in (42, 46) and complete<br />

- 23 -


Introduction<br />

response rates of greater than 75% for in-transit melanoma and greater than 80% for<br />

non-resectable extremity sarcoma were published (62-64).<br />

To complete this overview, it has to be mentioned that several other articles were<br />

published using mitomycin C, cisplatin and 5-fluorouracil in clinical protocols (54, 56,<br />

65), but the results obtained by these cytostatic drugs added no further advantages to<br />

the ones observed with melphalan.<br />

Table 3 summarizes recent publications of clinical investigations using melphalan<br />

and other chemotherapeutic drugs or combinatorial treatment regimens for cancer<br />

therapy. The role of TNF, its antitumor capacity and its clinical use alone or in<br />

combinatorial regimens will be discussed in the next chapter.<br />

- 24 -


Introduction<br />

- 25 -


Introduction<br />

1.6 TNF – heaven and hell in cancer therapy<br />

Tumor Necrosis Factor (TNF) was discovered by Carswell et al. in 1975. However,<br />

the trace of this molecule goes back through history far further. According to the writings<br />

of the Ebers Papyrus (c 1500 BC), the Egyptian physician Imhotep proposed ca. 2600<br />

BC to induce a local infection caused by a poultice and a small incision, as the<br />

recommended treatment for “swellings”, as tumors were described in those days. In<br />

Italy, the priest Peregrine Laziosi spontaneously healed from a huge malignant tumor<br />

after the tumor broke through the skin and became severely infected (reviewed in (66)).<br />

In Germany around 1740, the clinician Dr. Schwenke noticed that some patients with<br />

various malignancies could benefit from certain infectious disease progressions. The<br />

American surgeon W.B. Coley observed a complete remission of an egg-sized sarcoma<br />

after the surgical wound became infected with Streptococcus pyrogenes. Subsequently,<br />

he infected ten more cancer patients with these bacteria, resulting either in strong and<br />

finally fatal infection with a strong regression of malignancies, or in less to no<br />

inflammatory reaction and without tumor regression in the patients. Because of this<br />

unpredictability, Coley developed a treatment strategy using a bacterial “vaccine” of the<br />

killed bacteria S. Pyrogenes and Serratia marcescens, which was called “Coley’s<br />

toxins”. The success of these toxins was more or less limited to the tumor type<br />

(reviewed in (66, 67)).<br />

In contrast to clinical interest in the curative use of toxins, a wide range of microorganisms<br />

and microbial products have been studied as antitumor agents, including<br />

bacteria, fungi, plasmodia and trypanosomes. One of the most prominent bacteria was<br />

the mycobacterium Bacillus Calmette Guérin (BCG), which was shown to induce severe<br />

haemorrhagic necrosis of certain mouse tumors upon injection (68). The systemic<br />

antitumor effect of BCG (-extract) injection was thought to be due to a general increase<br />

in immunological reactivity, but the mechanism or the molecule of action was still<br />

unknown. The discovery of TNF by Carswell et al. provided a clue as to how these<br />

diverse reactions in response to microbial infections might be linked (68) (reviewed in<br />

(69)).<br />

After the initial publication a number of investigations turned to the study of TNF,<br />

although the enormous toxic side-effects in the patients discredited its systemic use for<br />

cancer patients. The intensive investigations on this molecule led to at least a complete<br />

description of the biological factor and were closely linked to the identification of<br />

bacterial structures responsible for the inflammatory response. Shear and colleagues<br />

have been the first who succeeded in purifiing a microbial protein present in the bacterial<br />

wall, which was termed bacterial polysaccharide, later on lipopolysaccharide (LPS) and<br />

seemed to be very effective in inducing tumor necrosis (reviewed in (70)).<br />

- 26 -


Introduction<br />

TNF is initially synthesised as a pro-hormone containing an aminoterminal peptide<br />

of varying length, depending on species. The propeptide segment of the molecule is<br />

highly conserved, showing an 86% homology between mouse and human. TNF consists<br />

of 157 amino acids with a molecular weight of 17 kD. In humans the molecule is nonglycosylated,<br />

in contrast to the murine form. In its biologically active form it consists as a<br />

non-covalently linked homotrimer. TNF can be found in an active membrane-bound<br />

precursor form and in a soluble form after cleavage from the membrane through a matrix<br />

metalloproteinase, the so-called TNF alpha converting enzyme (TACE) (71, 72). TNF<br />

was found to be produced mainly by monocytes and macrophages (73) upon activation<br />

by various stimuli, but also by neutrophils (74-77), basophils/mast cells (78, 79) and<br />

natural killer cells (80, 81). Further more, some other cell types have also been found to<br />

produce and secrete soluble TNF including astrocytes and glia cells in the nervous<br />

system and synovial cells in rheumatiod arthritis (82).<br />

Now, more than 30 years after its discovery, a whole spectrum of pleiotropic effects<br />

of the molecule is known, including physiological and pathophysiological actions. After<br />

the molecule has been cloned in the 1980’s independently by the groups of Fiers,<br />

Pennica and Shirai (83-86), TNF was shown to be a key mediator in inflammation and<br />

sepsis, immunological reactions, apoptosis, rheumatiod arthritis, liver regeneration and<br />

so on. (for further literature see (69, 70, 86-90). Between 1985 and 1988 recombinant<br />

TNF was made available to medical oncology and the hope for curing cancer raised.<br />

Unfortunately, systemic application in advanced cancer patients showed a very low<br />

maximally tolerated dose of 300 mg/m² (91) and a tumor response was hardly seen at<br />

that dose. Therefore, systemic application was more or less abandoned in 1988.<br />

1.7 A whole family of cytotoxic cytokines – the TNF superfamily<br />

Beginning with TNF, several other factors were identified to be able to induce<br />

death in multiple cell-types and organs in humans and other species. They all belong to<br />

a huge family, the TNF superfamily. The most prominent among them are TNF-ß (92),<br />

CD95L (FAS-L, Apo1L) (93) and TNF-related apoptosis-inducing ligand (TRAIL) (94).<br />

Interestingly, also the receptors of these molecules are closely related and belong all to<br />

a huge receptor family, the so-called TNF receptor superfamily. At present, more than<br />

19 members of the TNF superfamily are known that signal through 29 receptors (95).<br />

The identification of these cytokines and receptors was mainly driven by two<br />

scientific fields of interest, tumor-therapy and apoptosis. A detailed description of those<br />

- 27 -


Introduction<br />

ligands and receptors would be beyond the scope of this introduction. Therefore, only a<br />

few pairs of receptors and ligands will be highlighted which have been discussed<br />

intensively for tumor treatment.<br />

In 1968 Williams and Granger described a protein produced by lymphocytes<br />

capable of inducing death of tumor cells (92). After purification and determination of the<br />

amino acid sequence, the relationship to TNF was revealed (84, 96-100). The protein<br />

was first named TNF-ß, in view of the large homology to TNF, later lymphotoxin α<br />

(reviewed in (95)). By direct cloning strategies a number of further ligands closely related<br />

to TNF were identified (93, 99, 101). One among them was CD95L (Fas-L, Apo1L). A<br />

third ligand also being able to induce death of other cells was found independently by<br />

two different groups. Wiley and colleagues published a new type II membrane protein of<br />

281 and 291 aa which they named TRAIL and which in between has been shown to<br />

induce cell death in a broad variety of transformed cell lines (94). One year later in 1996,<br />

Pitti and co-workers published a protein of 281 amino acids found by expressed<br />

sequence tag, which seems closely related to Apo-1L and therefore named Apo-2L<br />

(102). By comparison of both proteins and their DNA-sequence it became clear that<br />

TRAIL and Apo-2L are identical.<br />

In general, all members of this TNF superfamily exert physiological and<br />

pathophysiological actions. Bharat B. Aggarwal therefore named their signaling<br />

pathways a double-edged sword (95). There is no doubt that the bright side of the TNF<br />

superfamily is their anticancer potential. This and the problems associated with this<br />

molecule have already been discussed.<br />

In the next sections two other members of this family will be highlighted, namely<br />

CD95 and TRAIL, which have been shown to be key players in physiology and<br />

pathophysiology.<br />

- 28 -


1.7.1 CD95, Fas and APO-1<br />

Introduction<br />

In 1989 two independent groups reported an unexpected induction of apoptosis by<br />

an antibody generated against tumor antigens (101, 103). Three years later the<br />

corresponding cytokine receptor could be cloned and was termed Apo-1 (104) by the<br />

Krammer group and Fas by the Nagata group (93). Finally, in 1994, the natural ligand<br />

CD95 was purified and cloned (105, 106).<br />

CD95 (Apo-1, Fas) was found to be expressed in many cell types, including heart,<br />

liver, lung and thymus (107). On the contrary, CD95L (Apo-1L, FasL) seems to be most<br />

likely expressed in the immune system, especially produced by T-cells and natural killer<br />

cells (105). In the following years, CD95/CD95L was found to play a major role in liver<br />

apoptosis (108-111), but similar to the other members of the TNF superfamily also<br />

physiologic functions of CD95 have been found, including removal of tumor cells (112),<br />

transducing growth promoting signals in proliferating T-cells (113, 114) and fibroblasts<br />

(115) and implication in liver regeneration after partial hepatectomy (116). Up-regulation<br />

of CD95 has also been demonstrated in active viral hepatitis (117). Since CD95L was<br />

found to be able to kill tumor cells, hopes were raised again for having a new anti-cancer<br />

drug. But this sword also turned out to be double-edged. Recombinant human CD95<br />

proved to be insufficient for general cancer therapy, because of severe liver toxicity<br />

observed in mice (118). In addition, Xiao-Zhong Wang and co-workers found that CD95<br />

and CD95-L were present in the majority of specimen collected from 50 HCC<br />

(hepatocellular carcinoma) patients. CD95 and CD95L were detectable in large numbers<br />

on carcinoma cells, indicating that these cells strongly express both (119). Similar<br />

findings have also been made by the group of Krammer et al. before (120). The question<br />

why these ligands are expressed raised a number of hypotheses. First, co-expression of<br />

receptor and ligand might lead to a so called fratricide death of HCC cells, induced not<br />

only by lymphocytes but also by their own ligands in an autocrine or paracrine manner.<br />

Such a mechanism might be involved in drug-induced HC apoptosis during tumor<br />

therapy. Evidence for this hypothesis is raised by the finding that bleomycin is able to<br />

upregulate CD95/CD95L in HepG2 cells (121). Second, CD95L might be important<br />

during the infiltration of tissue as well as during the dissemination into the liver (122), as<br />

demonstrated in hepatic metastasis of colon cancer cells (123). Finally, a third possibility<br />

is termed “tumor-counterattack”. This hypothesis implies a self-defence strategy of the<br />

tumor (cells). By expressing FasL on their surface, tumor cells will be able to induce<br />

apoptosis to antitumor immune cells instead of being killed by them and therefore<br />

survive. This was hypothesized by Stand et al. underlying the observation that HepG2<br />

cells, expressing CD95L after treatment with cytostatic drug, were able to kill CD95positive<br />

Jurcat t-cells. Furthermore, high prevalence of expression of CD95L in diverse<br />

- 29 -


Introduction<br />

tumor types (including colon cancer, liver cancer, melanoma and lung cancer) suggests<br />

that CD95L might be a general, perhaps an essential factor in the inhibition of anti-tumor<br />

responses (124). Since this hypothesis was formed, a hot debate has been started (for<br />

further literature see (125-127)]. Nevertheless, if this idea is true or wrong, CD95L or the<br />

agonistic antiCD95 antibody failed for cancer therapy.<br />

1.7.2 TRAIL<br />

Based on the homology of its extracellular domain to CD95L (28% identical), TNF<br />

(23% identical) and LTα (23% identical) (94) a new member of the TNF superfamily,<br />

TRAIL, was isolated in 1995 by Wiley and co-workers. Similar to others, TRAIL was<br />

found to be a type II transmembrane protein which is able to induce apoptosis in various<br />

transformed cell lines after binding to its specific receptor (94). Interestingly this factor<br />

did not seem to be toxic to normal cells in vitro. The identification of a number of<br />

receptors binding TRAIL, including DR4 (TRAIL-R1) (128), DR5 (TRAIL-R2, TRICK2)<br />

(128-130), DcR1 (TRID, TRAIL-R3) (128, 129, 131) and TRAIL-R4 (DcR2) (132, 133)<br />

(for review see also (134)) was confusing. Whereas DR4 and DR5 contain an<br />

intracellular death domain and are therefore able to induce apoptosis via DISC (death<br />

inducing signaling complex)-formation and subsequent caspase activation, DcR1 and<br />

DcR2 seem to function as decoy (Dc) receptors with no intracellular domains. Yet the<br />

function of these two receptors is still under debate. Corresponding transcripts of these<br />

receptors have been found in many human tissues, including foetal liver, adult testis, but<br />

not in most cancer cell lines examined (133, 134) (for review see (135)). Interestingly, it<br />

was found that TRAIL-Rs 1 and 2 seem to be able to induce two different types of cell<br />

death, apoptosis after intracellular recruitment of FADD (Fas associated death domain)<br />

followed by caspase-8 recruitment and activation (122, 136-138), or necrosis after<br />

recruitment of RIP (receptor-interacting protein) (139). In vivo studies performed by<br />

Walczak and co-workers revealed that neither murine TRAIL nor human TRAIL lead to<br />

general or organ specific toxicity or alterations in mice. In contrast, the authors could<br />

demonstrate that intravenous application of TRAIL led to a lowered tumor burden in<br />

SCID mice and the survival of tumor-injected mice was prolonged (140). Since it is<br />

known that TRAIL also induced apoptosis in human hepatocellular carcinoma and<br />

cholangiocarcinoma cells (141, 142), Gores and Kaufmann reviewed the question<br />

whether TRAIL is toxic to the liver from the view of hepatology and oncology (143).<br />

In their summary they state that current data suggest TRAIL to be the most<br />

promising cytokine for cancer therapy. But only less information is accumulated and<br />

- 30 -


Introduction<br />

reports stating that TRAIL might be toxic to normal hepatocytes should not be ignored<br />

(144, 145). All in all, TRAIL needs much more attention before early clinical trials might<br />

be started. As the authors argue, it is not known if pathophysiological perturbations or<br />

the influence of (cytostatic) drugs might render hepatocytes susceptible to TRAIL (for<br />

further discussion see also (145-155)). Therefore, more pre-clinical studies have to be<br />

undertaken.<br />

1.8 Combinatorial therapy<br />

Metastatic or primary non-resectable cancers confined to the liver, mostly resulting<br />

from metastatic colon carcinoma, still represent a significant clinical problem. Of 140.000<br />

new patients diagnosed with colorectal cancer in the US each year, approximately 20%<br />

to 30% will die of progressive metastatic disease confined to the liver (156-158). For<br />

Europe the prognosis is not better. In the UK colorectal cancer is the second most<br />

common cause of cancer death, with more than 40% of these patients destined to die of<br />

the disease despite current clinical management (159). More than 70% of all patients<br />

with colorectal cancer will develop liver metastasis during progression of the disease<br />

and autopsy studies indicate that in at least half of these cases the liver is the sole<br />

metastatic site (157). For those patients, a complete surgical resection has remained the<br />

only curative treatment strategy, with a five year survival rate of 40%. Only 10% to 20%<br />

of the patients are possible candidates for such surgery (46) and the vast majority of<br />

colorectal metastases confined to the liver are considered to be non-resectable (160,<br />

161). For the remaining 80%, prognosis is grim if the disease is confined to the liver: a<br />

mean survival rate of ten months (162) and less than six months with concomitant<br />

extrahepatic metastasis or regional recurrence (163, 164). Standard application of<br />

chemotherapeutic drugs prolongs this mean survival up to ten to fourteen months (165).<br />

Therefore, new methods, technologies or treatment strategies have been taken into<br />

account. But also other approaches like arterial embolization, chemoembolization,<br />

ethanol injection, radio therapy, cryosurgery and radiofrequency ablation did not result in<br />

prolonged survival (46). The most promising idea was an isolated setting with increased<br />

drug concentration at the tumor site using melphalan and TNF together.<br />

- 31 -


Introduction<br />

1.8.1 TNF – the good, the bad or the ugly in ILP<br />

Initial results using TNF and melphalan in isolated limb perfusion (ILP) achieved a<br />

complete tumor regression in 90% of the cases for patients with melanoma lesions (61).<br />

These and other preliminary results in the field of ILP with melphalan and TNF led to<br />

further clinical investigations on the effect of cytostatic drugs combined with cytokines<br />

also in isolated hepatic perfusion. However, the first results were disappointing. In a<br />

phase I clinical trial in 1994 using melphalan and cisplatin, tumor regression was<br />

registered in only 20% of treated patients (54). Another clinical study in the same year<br />

revealed also positive results for the use of TNF (166). In this article TNF was discussed<br />

to be mainly responsible for the breakdown of tumor vascularity. Therefore, TNF was<br />

also made responsible for the capillary leak which was often observed during IHP.<br />

In contrast, Alexander and co-workers could demonstrate that this augmented<br />

capillary leak during IHP occur via TNF-independent mechanisms (167). In this paper, a<br />

study of 27 patients was presented showing no leakage of hepatic perfusate,<br />

independent of TNF application. Furthermore, TNF did not affect melphalan tumor<br />

concentrations after IHP, too. In 1997 Borel Rinkes et al. published a study performed in<br />

pigs using TNF in the presence and absence of melphalan (168). In this study, which is<br />

one of the few pre-clinical trials, they found that IHP is technically feasible without<br />

systemic toxicity, mild transient hepatotoxicity, minimal systemic drug exposure and<br />

minor transient disturbances of liver biochemistry and histology.<br />

1.8.2 Investigation of hyperthermia in ILP<br />

In 1998, a number of clinical reports were published using melphalan and TNF<br />

concomitantly. De Vries et al. reported a partial response (PR) in five of six cases using<br />

1 mg/kg melphalan and 0.4 mg TNF with simultaneous hyperthermia of >41°C (58).<br />

However, also a mortality of 33% was stated. Hyperthermia in general has been shown<br />

to induce apoptosis in human gastric carcinoma cell lines in a p53-dependent manner<br />

(169). For Hela cells it could be shown that an intact p53 gene is essential for heatinduced<br />

apoptosis of cancer cells (170) and studies on human colorectal carcinoma cell<br />

lines revealed that p53-inactivated cells did not undergo apoptosis in response to<br />

hyperthermia, whereas normal cancer cells died (171).<br />

The use of hyperthermia in clinical protocols has been published by several<br />

authors. Hafström et al. reported a PR in three of eleven cases with a mortality of 18%<br />

- 32 -


Introduction<br />

(172) and Oldhafer et al. reported one complete response and 26 PR with 0% mortality<br />

using 60-140 mg melphalan and 200-300 µg of TNF (173). Last but not least, Alexander<br />

et al. published a phase II trial with an overall response rate of 75% (1 CR, 26 PR) using<br />

hyperthermia, 1.5 mg/kg melphalan and 1.0 mg of TNF (174).<br />

On the contrary to these promising results Lindner et al. reported a clinical trial in<br />

1999 in which eleven patients with non-resectable liver malignancies were treated with<br />

0.5 mg/kg melphalan and 30-200 µg TNF. In this study patients with metastases<br />

resulting from colon carcinoma displayed no response and only three patients with<br />

metastases from malignant melanoma showed partial response. In conclusion the<br />

authors stated that this regimen is a method with high toxicity (two patients died within<br />

the postoperative month). But it is not possible to distinguish between toxicity resulting<br />

from TNF and toxicity due to the operative procedure (65). On the other hand,<br />

melphalan/TNF was still successful in isolated limb perfusion as published by Plaat et al.<br />

1999 for the treatment of soft tissue sarcoma (175).<br />

Finally, in 2001 Alexander et al. published a study including 51 patients with nonresectable<br />

malignancies confined to the liver. Thirty-two of them were treated with<br />

melphalan/TNF, nineteen with melphalan alone. Despite one perioperative death (2%)<br />

patients displayed an overall objective radiographic response rate of 76% (38 of 50<br />

assessable patients) with a median duration of 10.5 months (range, 2 to 21 months).<br />

The median survival was 16 months, respectively. The authors could observe 18 PRs in<br />

26 patients (69%) whose prior therapy had failed.<br />

To sum this efforts up Weinreich and Alexander wrote in their review on<br />

transarterial perfusion of liver metastases in 2002 (42) that it is possible to perform safe<br />

IHP in patients with non-resectable cancers confined to the liver, with a low morbidity<br />

and mortality. It is possible to mediate clinically meaningful regression of refractory or<br />

advanced cancers, but they also remark that further research is needed and that the role<br />

of TNF during IHP still remains unclear.<br />

- 33 -


Introduction<br />

1.8.3 Pre-clinical results of melphalan in cancer therapy<br />

Since melphalan was first used in clinical trials beginning in the 1940s and most<br />

progress has been made in clinical trials using various nitrogen mustards and cytokines<br />

alone or in combination, very few data concerning pre-clinical in vitro studies on cells or<br />

in vivo studies in animals are available. Among present publications, most interest has<br />

been focused on the role of glutathione. In 1991, Skapek and co-workers observed an<br />

enhanced melphalan-mediated toxicity in nude mice after buthionine sulfoximine (BSO)<br />

pre-treatment (176). BSO is a potent inhibitor of glutathione γ-glutamylcysteine<br />

synthetase, resulting in intrahepatic GSH depletion (88%). The toxicity included<br />

reversible gastrointestinal toxicity, myelosuppression and histologic evidence for<br />

nephrotoxicity, but no evidence for hepatic failure. In 1992, Laskowitz et al. reported<br />

enhanced melphalan-induced gastrointestinal toxicity after regional hyperthermia and<br />

BSO-mediated GSH depletion (177). In 1999, Smith and co-workers investigated the<br />

effect of BSO on melphalan-induced toxicity in mice. The authors found that a low dose<br />

intravenously BSO administration did not alter melphalan toxicity (5mg/kg used),<br />

whereas a high dose application lead to potentiated myelotoxicity, bone marrow toxicity<br />

and nephrotoxicity (178). Interestingly, no liver toxicity was assessed, although Kramer<br />

et al. demonstrated enhanced liver toxicity with BSO and alkylating agents in Fischer<br />

rats before (179). In 1999 Vahrmeijer and colleagues investigated the effect of GSH<br />

depletion on inhibition of cell cycle progression and the induction of apoptosis by<br />

melphalan in human colorectal cancer cells (180). In this manuscript, the scientists<br />

reported that GSH depletion below 20% of normal cellular GSH level enhanced<br />

melphalan-induced cytotoxicity nearly 3-fold and DNA fragmentation was 4-fold higher<br />

than compared to the non-GSH depleted state, indicating enhanced apoptosis. Further<br />

analysis revealed that mitochondrial membrane potential was also enhanced both in<br />

melphalan and melphalan/BSO treatment.<br />

In summary, numerous cellular studies had been undertaken to reveal GSX<br />

dependent changes in melphalan-induced cytotoxicity resulting in enhanced cellular<br />

toxicity. This might be due to the fact that melphalan is able to react spontaneously with<br />

GSH independent of the activity or presence of glutathione-S-transferase (181). Studies<br />

performed in man and rat by Vahrmeijer and colleagues confirmed these results<br />

indicating that hepatic GSH conjugation plays a very minor (if any) role in the elimination<br />

of melphalan (182). It might be possible, especially in low dose therapy that the more<br />

GSH is present in the cell, the more alkylating mustard is eliminated. On the contrary,<br />

when GSX is depleted more molecules of alkylating agents might be able to crosslink<br />

cellular DNA and subsequently toxicity is enhanced.<br />

- 34 -


1.9 Apoptosis<br />

Introduction<br />

Today’s knowledge in the field of apoptosis mainly originates from research<br />

activities in the second half of the 20 th century, although the German pathologists<br />

Virchow (183) and Weigert have been interested in cell death nearly one century before.<br />

They termed cell death “necrosis” or “coagulative necrosis” (Weigert 1877), derived from<br />

the Greek expression for death.<br />

Beginning with the second half of the 20 th century, developmental biologists<br />

recognised that there must be different forms of cell death (184), in view of<br />

morphological changes and the shape of death. Additionally a defined pattern seems to<br />

regulate these forms of cell death, since only defined cell populations have been<br />

destined to die. Investigation of the development of the worm Caenorhabditis elegans<br />

revealed that exactly 131 of the 1,090 cells die during development (185, 186). Because<br />

this cell death seems to be part of a developmental plan or program, the term<br />

“programmed cell death” was used for this shape of death. Finally Kerr, Willy and Currie<br />

defined this kind of death as “apoptosis”, based on the morphologic characteristics of<br />

this type of death (187). It could be clearly distinguished from necrosis which represents<br />

a pathological state of a cell (organ), including loss of the integrity of the plasma<br />

membrane, disruption of the membrane potential and a spill-out of cellular contents in<br />

the environment (188, 189) finally resulting in inflammation.<br />

Apoptosis is evolutionary conserved and can be found in many organisms ranging<br />

from plants and insects to birds and mammals (for review see (190). Apoptosis is<br />

necessary for foetal and embryonal development, responsible for a proper function of<br />

the immune system and for tissue homeostasis. But there is also recent evidence that<br />

dysregulation of apoptosis is involved in a number of human diseases, including AIDS<br />

(acquired immunodeficiency syndrom), viral infections, neurodegenerative disorders<br />

(Alzheimer’s disease, Parkinson’s disease), autoimmune diseases, ischemic disorders<br />

(myocardial infarction, stroke, reperfusions injury), toxin-induced liver diseases and<br />

cancer (191). In view of the huge information material according to the knowledge about<br />

induction and regulation of apoptosis, the biochemical processes and the involvement in<br />

physiological and pathophysiological processes, it is not possible to discuss this topic in<br />

depth here, because this would be beyond the scope of this introduction. Many of these<br />

topics have been covered by a number of excellent reviews, which I would like to refer to<br />

(189, 192-210). Only the role of apoptosis in cancer, the differences of apoptosis in<br />

primary versus transformed cells and signaling of TNF receptors in apoptosis will be<br />

discussed very shortly because of its importance for the current topic.<br />

- 35 -


Introduction<br />

1.9.1 Signaling of TNF receptors in apoptosis and anti-apoptosis<br />

Since TNF was introduced as a molecule with a whole plethora of actions, with<br />

great importance in physiology and pathology, it is also important for this thesis to give a<br />

short overview about the signaling of TNF and TNF receptors, in order to understand<br />

how these pleiotropic effects are transmitted to the cell.<br />

TNF mediates its cellular effects mainly by two distinct surface receptors, namely<br />

the 55 kD TNFR1 and the 75 kD TNFR2 (211, 212). Interaction between TNF and<br />

TNFR1 activates numerous signal transduction pathways in multiple subcellular<br />

compartments including the plasma membrane, cytosol, endosomes, mitochondria and<br />

the nucleus (211). Among these diverse cellular effects of this pleiotropic cytokine, two<br />

effects have gained the most attention: the initiation of cell death and the activation of<br />

transcription factors, leading to the expression of genes, some of which promote cellular<br />

survival, some of which are involved in cell growth, development, oncogenesis and<br />

some of which are important in immune, inflammatory and stress responses. The<br />

balance of these pathways at least determines the cellular fate.<br />

To date, most of the players in the TNF pathway have been validated. However,<br />

many questions remain poorly understood. For example, TNF is known to induce<br />

necrotic and apoptotic cell death (213). Which signaling molecules or cellular conditions<br />

decide which of these morphologically separated forms of cell death are induced?<br />

Furthermore, the crosstalk between apoptosis and the pathways leading to cellular<br />

proliferation including NFκB and JNK signaling pathways is still not clear and has to be<br />

further investigated.<br />

- 36 -


1.9.2 The role of TNF receptor 1<br />

Introduction<br />

A common feature of each pathway is the TNF-induced formation of an intracellular<br />

signaling complex at the cell membrane, in case of cell death this is called deathinducing<br />

signaling complex (DISC). The demise of a cell can occur by two distinct<br />

mechanisms, apoptosis and necrosis and TNF can elicit both of them. In most cell types,<br />

induction of apoptosis by TNF cannot occur without inhibiting the expression of new<br />

genes (214) indicating that TNF might also induce cellular signals independent of cell<br />

death via binding to their specific receptors. Indeed, TNF has been shown to activate<br />

survival pathways inhibiting activation of death pathways (215). The pathway leading to<br />

the activation of caspases (216) and finally to apoptotic cell death is primarily mediated<br />

by TNFR1 (217), because the cytoplasmatic domain of TNFR1 contains a region called<br />

death domain which has been shown to be essential for TNF-induced cytotoxicity (218).<br />

In the basal state without bound TNF the death domain (DD) is associated to a silencer<br />

(SODD) which prevents signal transduction (219). Upon binding of trimeric TNF to the<br />

extracellular domain of TNFR1 the inhibitory protein SODD is released from the<br />

intracellular death domain (219, 220) and aggregation of TNFR1 death domains is<br />

enabled. The aggregated receptor is recognised by the adapter protein TNF receptorassociated<br />

death domain (TRADD) which exhibits high affinity for the aggregated death<br />

domains (221, 222) and forms a stable complex with the intracellular domains. The<br />

death domain of TRADD further interacts with the DD of FADD (Fas-associated death<br />

domain), forming also a stable complex (223, 224). To stabilise this huge protein<br />

complex, further proteins are required (222). FADD contains two major domains, the DD<br />

and a N-terminal domain termed death effector domain (DED) (223, 224). The<br />

interaction of TRADD and FADD leads to an altered conformation of the protein with an<br />

exposition of the death effector domain (225) which has an important function. It is<br />

analogous to a so-called caspase- activating domain (CARD) found in the prodomain of<br />

various upstream caspases including procaspase-8 (226). Upon recruitment to the DISC<br />

procaspase-8 becomes activated autoproteolytically (227-229). Next to procaspase-8<br />

procaspase-10 is also be recruited and activated at the DISC (230). Activated caspase-8<br />

on the one hand is able to activate the effector proteases caspase-3, -6 and -7, which in<br />

turn cleave multiple cellular proteins, resulting in cell death (230).<br />

On the other hand, activation of effector caspases is amplified by a mitochondrial<br />

pathway (231). A slight caspase-8 activity is sufficient to activate Bid, a Bcl-2 family<br />

member which then translocates to the mitochondria,. Bax, another Bcl-2 family<br />

member, is activated by this pathway to localise to the mitochondrial membrane and<br />

increases its permeability resulting in mitochondrial dysfunction (reviewed in (199)) and<br />

release of cytochrome c and other mitochondrial proteins (232-236) Cytochrome c binds<br />

- 37 -


Introduction<br />

to Apaf-1 and pro-caspase-9 connected via a CARD-domain and forms a protein<br />

complex, the so-called apoptosome, of which dATP is an integral part and thus regulates<br />

the apoptotic signaling via subsequent autocatalytic activation of pro-caspase-9 (231,<br />

237, 238). Active caspase-9 in turn is able to activate further downstream caspases<br />

(199, 239, 240). In addition, another factor called apoptosis inducing factor (AIF) is also<br />

released from the mitochondria and translocates to the nucleus to effect degradation of<br />

the DNA into large fragments (241).<br />

1.9.3 The role of TNF receptor 2<br />

TNF binding to TNF-receptor 1 has been shown to be the major ligand-receptor<br />

interaction involved in apoptotic signal transduction. In contrary to TNFR1, TNFR2 does<br />

not posses an intracellular death domain and hence, is unable to recruit death adapter<br />

molecules like TRADD or FADD. Under certain circumstances, however, TNFR2 may<br />

also be involved in inducing apoptosis (242-245). For example, TNFR2 seems to be the<br />

major player in apoptosis of mature CD8 + T-cells (246) and antagonistic antibodies<br />

against TNFR2 are able to inhibit, at least partially, TNF-induced cytotoxicity of tumor<br />

cells (247). Evidence for involvement of TNFR2 in liver injury comes from studies of Con<br />

A-induced liver toxicity in mice, since mice lacking TNFR2 are fully protected from Con<br />

A-induced liver failure (248, 249). This suggests that both receptors upon stimulation<br />

might contribute to cellular reactions leading to cell death. However, the use of<br />

antagonistic antibodies in high concentrations and molecular receptor analysis revealed<br />

that only TNFR1 is responsible for cytotoxic activity (242, 250-254). How can this be<br />

explained? Two roles for TNFR2 in apoptotic signaling have been postulated. First,<br />

TNFR2 might increase the local concentration of TNF so as to enhance death signaling<br />

from TNFR1. Evidence for this hypothesis comes from the fact that the dissociation rate<br />

of TNF from TNFR2 (t1/2= 10 min) is significantly greater than that for TNFR1 (t1/2> 3 h)<br />

whereas TNFR2 has a higher affinity to TNF than TNFR1 (Kd 100pM vs. 500pM) (255).<br />

Therefore, TNFR2 preferentially binds TNF at low concentrations (256). Hence, TNFR2<br />

can perform a ligand passing function and binds many molecules of TNF, which are<br />

rapidly passed to TNFR1 (255, 257, 258). Whether this passing is simply a result of<br />

TNFR2-dependent increase in local ligand concentration or whether the mechanism<br />

requires formation of transient heterocomplex is not clear yet. Second, activation of<br />

TNFR2 might induce ubiquitinylation and subsequent degradation of TRAF2, which<br />

inhibits the formation of the DISC by TNFR1 and which activates anti-apoptotic<br />

processes induced by TNF (259, Duckett, 1997 #1280, 260). Evidence for this fact<br />

comes also from observations of Fotin-Mleczek et al. demonstrating that TNFR2 induces<br />

- 38 -


Introduction<br />

depletion of TRAF2 and inhibitor of apoptosis proteins (IAPs) and therefore accelerates<br />

TNFR1-dependent activation of caspase-8 (261). To get a better understanding of these<br />

mechanisms induced by TNF, the signaling of TNF receptors will be shortly discussed in<br />

the next chapter.<br />

1.9.4 Anti-apoptotic signaling<br />

The activation of transcription factors is a major cellular effect induced by TNF<br />

(reviewed in (211)). In general intracellular pathways specific for TNFR1 or specific for<br />

TNFR2 have been described, as well as signaling pathways shared by both receptors.<br />

NFκB activation is associated with TNF-induced cellular protection. Both TNF<br />

receptors are able to activate NFκB independently (221, 262, 263). For TNFR1 the<br />

protein RIP is essential as scaffold for TRAF1 and TRAF2 heterodimers on the contrary<br />

to TNFR2, which can directly associate with TRAF1-TRAF2 heterodimers (262). TRAF2<br />

has been shown to interact with NFκB-inducing kinase (NIK) (264-266) which is part of<br />

the high molecular weight complex IκB-kinase (IKK also referred to as signalosome)<br />

(266-270). Activation of NFκB relies on phosphorylation-dependent ubiquitinylation and<br />

degradation of inhibitor of κB (IκB) proteins, which normally retain NFκB within the<br />

cytoplasm (271, 272). The phosphorylation of IκB is mediated by IκB-kinase (273). IκB is<br />

phosphorylated and subsequently degrades from NFκB. NFκB translocates to the<br />

nucleus and activates transcription. Alternatively, NFκB can be activated specific via<br />

TNFR1 upon recruitment of IRAK (interleukin-1 receptor-associated kinase) and<br />

subsequent formation of a complex between IRAK and NIK (274-276).<br />

c-Jun NH2-terminal kinase (JNK) activation is also associated with TRAF2 and<br />

both TNF receptors have been shown to be initiate JNK activation (262, 277). Briefly,<br />

TRAF2 interacts with three proteins; TRAF-associated NF-kB activator (TANK) activates<br />

GCK(R) which in turn activates mitogen-activated protein kinase/extracellular signalregulated<br />

kinase kinase kinase 1 (MEKK1). MEKK1 has been shown to be a potent<br />

activator of the JNK pathway (277-279). Active JNK can phosphorylate and activate<br />

various transcription factors including c-Jun, ATF-2, Ets-like protein 1 (Elk-1) and cAMPresponsive<br />

element binding protein (CREB) (277, 279, 280). Alternatively, JNK can also<br />

be activated specifically by TNFR1 involving the activation of phosphatidylinositol 3<br />

(PI3)-kinase via the downstream mediator Rac and cytosolic phospholipase A2.<br />

Additionally, there is also evidence that TRAF2 is initiating the recruitment of cellular<br />

inhibitor of apoptosis proteins (cIAP) (281). IAPs are capable to inhibit caspase activity<br />

- 39 -


Introduction<br />

directly (282-284). These inhibitory proteins contain two domains, baculoviral inhibitory<br />

repeat (BIR)-1 and BIR-2, which are able to block activity of the effector caspases-3, -6<br />

and -7, whereas BIR-3, a third domain of cIAP, selectively inhibits caspase-9 activity<br />

(284).<br />

A third pathway onset by both TNF receptors is p38 kinase activation. Following<br />

TRAF-2 recruitment apoptosis signal-regulating kinase 1 (ASK1) is known to bind TRAF-<br />

2 (285). ASK1 itself can directly activate three MAPKKs (MKK2, -3, -6) (286-288).<br />

Substrates for p38 kinase include cytosolic phospholipase A2 and activating transcription<br />

factor 2 (ATF2), both substrates assists in activator protein 1 (AP-1) transcription factor<br />

formation (289, 290).<br />

To complete this discussion, three more pathways have been shown to be<br />

activated by TNFR1: the ERK kinase activating pathways, neutral sphingomyelinase<br />

pathway and acid sphingomyelinase pathway (reviewed in (211)). The last one links<br />

TNF to lysosomal protease cathepsin D, thereby contributing to TNF-induced cell death<br />

(291). This is of greater interest since this family of proteases has been discussed to be<br />

involved in caspase-independent apoptosis, a phenomenon that attracts more and more<br />

attention.<br />

1.9.5 Modulation of apoptosis by changing cellular redox and energy state<br />

Glutathione<br />

The tripeptide glutathione (GSH, γ-glutamyl-cysteinyl-glycine) represents the most<br />

abundant non-protein thiol in the cell, serving as the major antioxidant and providing<br />

defence against xenobiotics as a phase II conjugation substrate (292-294). Glutathione<br />

affects numerous central cellular functions, like DNA and protein synthesis, gene<br />

transcription, transport, catalysis, cell growth and apoptosis (203, 295, 296).<br />

Interestingly, modulation of intracellular hepatic glutathione has been shown to<br />

protect mice from death receptor-induced liver injury mediated by CD95 (297). In this<br />

study the underlying mechanism for this apoptosis protection by glutathione depletion<br />

was investigated and caspase-3 activity was identified as the major site being<br />

dependent on the presence of GSH, while GSSG attenuated the proteolytic activity<br />

(297). Furthermore, it could be demonstrated that not only apoptosis but also necrosis<br />

occurred following death receptor activation by TNF in the D-galactosamine (GalN) /TNF<br />

- 40 -


Introduction<br />

and GalN/LPS-model, as well as in LPS-shock and ConA-induced liver damage. Cell<br />

death induced by acetaminophen was enhanced under glutathione-depleting conditions<br />

(298), which might be due to generation of the toxic metabolite N-acetyl-pbenzoquinoneimine,<br />

leading itself to depletion of intracellular glutathione, alteration of<br />

redox potential and ultimately, cellular necrosis (299).<br />

All these studies have been undertaken in healthy animals and the in vivo effects<br />

could not be demonstrated in vitro in isolated primary hepatocytes, therefore implications<br />

for cancer have been not investigated. For transformed cells it could be demonstrated<br />

that glutathione depletion mediated by IPTG or L-buthionine-(300)-sulfoximine (BSO), a<br />

specific inhibitor of GSH synthetase) induced severe apoptosis of Ha-ras-transformed<br />

NIH3T3 cells, mostly mediated through generation of reactive oxygen species (301)<br />

whereas for immortalised human keratinocytes a protection from UVA-induced apoptosis<br />

could be shown by BSO-mediated intracellular glutathione depletion (302). For<br />

cholangiocytes Celli et al. could demonstrate that glutathione depletion by BSO is<br />

associated with decreased (87%) levels of anti-apoptotic protein Bcl-2 and subsequently<br />

enhanced apoptosis (303). Last but not least, arachidonic acid-induced apoptosis was<br />

mostly mediated by generation of reactive oxygen intermediates and subsequent lipid<br />

peroxidation (304).<br />

All in all, glutathione depletion could be shown to protect mice from death receptormediated<br />

apoptosis whereas less information is available how glutathione depletion<br />

affects apoptosis in cell lines or cancer cells. The few data available mostly demonstrate<br />

enhanced apoptosis which is mostly due to enhanced radical formation subsequently<br />

inducing cell death.<br />

ATP<br />

Next to the redox state of the cell, apoptosis has also been shown to be influenced<br />

by cellular ATP levels. Investigations on death receptor-mediated apoptosis under ATPdepleting<br />

conditions revealed that reduced TNF-mediated apoptosis in contrast to<br />

enhanced CD95-mediated apoptosis (305). These effects could also be demonstrated in<br />

animal experiments (306) while the important ATP-dependent step in the apoptotic<br />

cascade is still not known. Infusions of D-fructose as a source of energy supply during<br />

parenteral nutrition have been used for many years and especially in patients with<br />

chronic or acute liver diseases fructose has been claimed to have several advantages<br />

over glucose (307, 308). The beneficial action of fructose had been assumed from the<br />

- 41 -


Introduction<br />

biochemical findings that this hexose is metabolised more rapidly than glucose (309).<br />

Fructose metabolism has been shown to be independent of insulin (307) and in liver<br />

diseases uptake of fructose by the liver is less affected than uptake of glucose.<br />

Interestingly, since the ATP-depleting capacity is closely related to the presence and<br />

activity of the rate-limiting enzyme fructokinase, investigations of Adelman, Morris and<br />

Weinhouse on the activity of fructokinase, triokinase and aldolases in liver tumors of the<br />

rat revealed that in these tumors hardly no activity of fructokinase could be detected<br />

(310). Indeed, most cell lines failed for ATP depletion by carbohydrates. On the other<br />

hand, isolated primary cells incubated with ATP-depleting sugars as fructose or tagatose<br />

and isolated livers of rat (309) or man (311) perfused with fructose can be ATP-depleted<br />

without toxic effects on the cells or the isolated organ. This discrepancy between primary<br />

versus transformed tumor cells is of further interest and might be useful also in<br />

chemotherapy. But currently no data on ATP depletion in clinical investigations of<br />

cancers confined to the liver are present. Despite this alteration in primary versus<br />

transformed hepatocytes a number of differences have been found in various cell types.<br />

These alterations will be shortly discussed in the following chapter.<br />

1.9.6 Differences in apoptosis of primary versus transformed cells<br />

As mentioned above, apoptosis is a generally controlled mechanism of cell death<br />

which is common to various cell types and organs and is evolutionary conserved<br />

throughout multicellular organisms. Tumor cells or immortalised cell lines instead display<br />

a number of differences in this control panel, leading to an altered response to death<br />

stimuli, including sensitivity to death signals and altered apoptotic pathways. Especially<br />

the death-inducing members of the TNF receptor superfamily have been studied<br />

intensively leading to the elucidation of their role in tumor evasion from the immune<br />

system (reviewed in (193)).<br />

One of the most prominent members of this family is TRAIL, which kills tumor cells<br />

very effectively but has no effect on healthy primary cells (94). Since at least five<br />

different receptors for TRAIL have been identified, it has been suggested that the<br />

differential expression of non-death receptors for TRAIL (TRAIL-R3) might be<br />

responsible for the different sensitivity to TRAIL observed for normal and transformed<br />

cells (129, 312). On the contrary, Leverkus et al. could demonstrate that TRAIL is able to<br />

induce apoptosis not only in transformed keratinocytes but also in primary keratinocytes<br />

in a dose-dependent manner, displaying a marked shift in sensitivity (313). This shift<br />

could be explained by intracellular modulation of adapter proteins and seems not to be<br />

- 42 -


Introduction<br />

related to the expression profile of TRAIL-receptors on the surface of the cell. Especially<br />

FADD and TRADD have been shown to modulate intracellular apoptotic signaling since<br />

expression of dominant negative forms of those proteins abrogate TRAIL-mediated<br />

apoptosis (130, 228). Further more, cFLIPL, a protein known to interfere with the<br />

proximal death receptor-mediated cascade, has been shown to suppress CD95-, TNFand<br />

TRAIL-mediated apoptosis (314-316) and was shown to be strongly expressed in<br />

primary keratinocytes, while it was undetectable in transformed cells (313). These<br />

findings might also be in line with previous publications on FADD since FLIP might<br />

compete with FADD for binding to pro-caspase-8 (317).<br />

Another publication identified the PI-3-kinase/AKT-pathway responsible for the<br />

major differences between primary and transformed keratinocytes in UV-induced<br />

apoptosis (123). On the DNA-level changes were most frequently observed in the tumorsuppressor<br />

p53-gene. Studies on apoptotic signaling in normal and neoplastic germinal<br />

center revealed that a number of intracellular proteins modelling apoptosis are<br />

dysregulated including Bcl-2 upregulation via chromosomal translocation of the bcl-2<br />

gene (reviewed in (318)). API2/MLT and Bcl-10 were found to be enhanced via<br />

chromosomal translocation in mucosa-associated lymphoid tissue (MALT) lymphoma<br />

resulting in a fusion protein of API 2 and MLT leading to NFκB-mediated inhibition of<br />

apoptosis (319, 320). Finally, Bax was found to be downregulated because of mutations<br />

leading to inactivation of the bax gene resulting in the loss of pro-apoptotic Bax-protein<br />

identified in a Burkitt lymphoma being resistant to CD95-mediated apoptosis (321).<br />

Independent of gene alterations, reduced levels of Bax were also found in several<br />

lymphoma malignancies (322, 323).<br />

On the receptor level mutations of CD95 were found to affect the intracellular death<br />

domain resulting in loss of DISC-assembling (324) and subsequently inhibition of<br />

docking and activation of pro-caspase-8 (325-328). Last but not least, downstream of<br />

receptor signaling, caspase-10 was found to be a common site of mutations in about<br />

15% of B-cell non Hodgkin’s lymphomas. Mutations in the death effector domain (DED)<br />

of the protein led to a loss of function of this domain which is necessary for interaction<br />

with FADD at the receptor site (329, 330).<br />

In summary, mutations resulting in altered apoptotic machinery could be identified<br />

at nearly all levels of intracellular apoptotic signaling making tumors less susceptible for<br />

killing by extracellular triggering. On the other hand the knowledge of these differences<br />

between primary and transformed cells represent new potential targets for molecular<br />

therapy or drug development raising hopes to find the tumors Achilles’ heel.<br />

- 43 -


1.9.7 Apoptosis in cancer<br />

Introduction<br />

Apoptosis plays a crucial role in maintenance of tissue homeostasis and serves as<br />

a safeguard system to prevent cell transformation through elimination of aberrant cells<br />

with dysregulated gene expression caused by mutation or amplification of DNA<br />

damages resulting from various physical, chemical or biological noxa which exceed the<br />

limit of repair (331). In this case, cells incurring an oncogenic mutation or integration of<br />

oncoviral genomes might be eliminated through the onset of the apoptotic program,<br />

either triggered by intrinsic mediators or by engagement of death receptors via their<br />

specific ligands presented by immune cells (332). Especially in hepatic carcinogenesis<br />

and HCV or HBV infections large clonal expansion of cytotoxic T cells is induced, which<br />

are able to induce CD95-mediated death in transformed or infected hepatocytes.<br />

Recent studies have shown that various anti-cancer drugs such as cisplatin,<br />

etoposide and 5-fluorouracil (5-FU) are able to induce apoptosis in cancer cells<br />

(reviewed in (333)), but a functional p53 gene is required (334, 335). Therefore,<br />

mutations in p53 have been shown to be concerned in no less than 50-60% of cancers<br />

in various types and therefore enabling cells to acquire resistance to apoptosis (331,<br />

336). Another gene isolated in rodents is an intronless member of the myc family called<br />

s-myc (337, 338). The resulting gene product has been show to induce apoptosis in a<br />

p53-independent manner when transfected in rat and human glioma cells (339-341). If<br />

the assertion that anti-cancer drugs are able to induce apoptosis in transformed cells is<br />

correct and due to the fact that apoptosis is a highly regulated biochemical process<br />

(reviewed in (196)) biochemical alterations might exist that make cells more or less<br />

susceptible or resistant to apoptosis. This might contribute to cellular or molecular<br />

resistance to a wide variety of drugs (342). Because of these potential implications for<br />

mechanisms of resistance, there is considerable interest in determining whether<br />

anticancer drug initiate the apoptotic process through death receptors or intrinsically via<br />

the mitochondrial pathway.<br />

One hypothesis which is currently discussed is the activation of death via CD95<br />

(Fas, Apo-1) triggered by the expression of CD95L on immune cells by anticancer drugs<br />

of various types including doxorubicin (DNA intercalating), etoposide (topoisomerase II<br />

inhibitor), cisplatin (DNA-crosslinking) and bleomycin (DNA-fragmentation) (121, 343-<br />

348). Nearly all of them have also been shown to upregulate CD95L mRNA (for review<br />

see (342)) due to drug induced activation of the transcription factors AP-1 and NFkB<br />

(101). Unfortunately, CD95L is not only upregulated in cytotoxic B and T cells, but also<br />

in target cells (cancer cells) as discussed earlier (124, 349, 350). Responsible for the<br />

expression of CD95L mRNA might not only be the activation of transcription factors, but<br />

- 44 -


Introduction<br />

also reactive oxygen intermediates which might contribute to CD95L upregulation as<br />

suggested by Hug et al. (351). In addition to CD95, other members of the TNF<br />

superfamily have been shown to be upregulated by anticancer drugs, indicating that<br />

these proteins might also play a role in apoptosis induction during cancer development<br />

(352).<br />

But does this really hold true? In their review on induction of apoptosis by cancer<br />

chemotherapy Kaufmann and Earnshaw discussed these facts very critical (342).<br />

Following their argumentation drug concentrations used to demonstrate CD95L<br />

upregulation in cancer cells are very high exceeding doses used in clinical approaches.<br />

Only for one drug (5-FU) it was demonstrated that CD95L is upregulated by drug<br />

concentrations having similar time courses and dose response curves for CD95L<br />

induction and inhibition of clonogenic survival (353, 354). Additionally, manipulations<br />

inhibiting CD95 and CD95L interactions have been shown to have opposite effects in<br />

different laboratories demonstrating no crucial role for CD95 (for review see (342)). Last<br />

but not least CD95-negative cells have been shown to react normally upon anti-cancer<br />

drug treatment (332, 355-357).<br />

Due to these conflicting results, it is of great importance to shed more light on the<br />

role of apoptosis in cancer, especially since mechanisms of drug resistance are not well<br />

understood, while displaying an enormous clinical problem. Finally, it is important to<br />

keep in mind that apoptosis induction via anti-cancer agents is not restricted to<br />

transformed cells, but also to normal healthy cells which should be prevented from<br />

apoptosis induction. Any knowledge that leads to differences in death induction between<br />

healthy and transformed cells might be extremely important and therefore can contribute<br />

to further advances in tumor therapy.<br />

- 45 -


2 Objectives of the thesis<br />

Objectives<br />

Liver metastases resulting from colon carcinoma or other types of cancer still represent<br />

a major clinical problem. Surgery is often not possible due to numerous metastases<br />

spread over the whole organ. Radiation results in massive necrosis leading to<br />

inflammatory liver failure and chemotherapy, even when applied locally, is either<br />

insufficient or leads to massive cell death also involving healthy tissue. The<br />

understanding of mechanisms leading to liver-specific tissue destruction caused by<br />

chemotherapeutic drugs, the role of cytokines and the interaction of different cell types<br />

being part of this scenario can therefore provide new rationale for therapeutic<br />

intervention strategies in cancer therapy. The cellular ATP content has been shown to<br />

be a central metabolic parameter modifying cell death in different cell types. A possibility<br />

for selective ATP depletion in hepatocytes is exposure to high concentrations of fructose<br />

or other sugars that rapidly trap intracellular phosphate. This approach has the<br />

advantage that sufficient residual ATP (>15% of control) remains in the depleted<br />

hepatocytes, thereby avoiding induction of necrosis due to total ATP loss. However, this<br />

metabolic ATP trapping has been shown to fail for most cancer cell lines providing an<br />

interesting tool for cancer therapy. At present the role of ATP depletion in cancer therapy<br />

has been poorly investigated and pre-clinical studies involving cell culture and animal<br />

models are hardly available.<br />

In the present study the mechanism of melphalan-induced hepatocyte toxicity has been<br />

investigated in primary liver cells and in isolated liver perfusion. The effect of ATP<br />

depletion in melphalan-induced liver cell-death was studied in liver cells and organs of<br />

healthy mice and in livers isolated from tumor-bearing mice.<br />

In particular, the following questions were addressed in the present thesis:<br />

1. What is the mechanism of melphalan-induced hepatotoxicity in<br />

isolated primary murine liver cells?<br />

2. Can this cell death be blocked by modification of the apoptotic<br />

machinery?<br />

3. How does ATP depletion affect melphalan-induced hepatotoxicity?<br />

4. Is there a difference in hepatic apoptosis induced by melphalan<br />

alone or in presence of exogenous soluble TNF?<br />

5. Is the isolated perfused mouse liver a sufficient model to reflect<br />

chemotherapy-induced liver failure in the murine system?<br />

6. How does ATP depletion act on isolated perfused livers of tumorbearing<br />

mice treated with melphalan and exogenous TNF?<br />

- 46 -


3 Materials and methods<br />

Materials and animals<br />

3.1 Chemicals<br />

Materials & Methods<br />

N-acetyl-Asp-Glu-Val-Asp-7-amino-4-trifluoromethylcoumarin (DEVD-afc) and<br />

Pefabloc® were obtained from Biomol (Hamburg, Germany). The caspase inhibitor<br />

benzoyloxycarbonyl-val-ala-asp-fluoro-methylketone (zVAD-fmk) was purchased from<br />

Alexis Biochemicals (Lausanne, Switzerland), as well as the recombinant Apo-2L Killer<br />

TRAIL. D-galactosamine (GalN) and Hepes was purchased from Roth (Karlsruhe,<br />

Germany) and LPS (Salmonella abortus equi) from Metalon (Wusterhausen, Germany).<br />

Melphalan was obtained from Glaxo Wellcome GmbH &Co (Bad Oldesloe, Germany)<br />

and Percoll was obtained from Amersham Biosciences (Freiburg, Germany),<br />

respectively. AlamarBlue was obtained from BioSource (Solingen, Germany).<br />

Cytotoxicity detection kit (LDH) was used from Roche Diagnostics GmbH (Mannheim,<br />

Germany). Concanavalin A (Con A), staphylococcal enterotoxin B (SEB), ketohexoses<br />

such as fructose, tagatose, mannose, mannitol, actinomycin D, DMSO and all other<br />

reagents and recombinant enzymes not further specified were obtained from Sigma<br />

(Deisenhofen, Germany). Pentobarbital (Narcoren®) was purchased from Sanofi<br />

Withrop (München, Germany).<br />

3.2 Antibodies and recombinant proteins<br />

Activating anti-CD95 antibody (Jo2) and polyclonal IgG-horseradish peroxidasecoupled<br />

secondary antibody (goat anti-mouse) were purchased from PharMingen (San<br />

Diego, CA, USA). Recombinant mouse TNF was obtained from Innogenetics (Ghent,<br />

Belgium). Antibody pairs (specific rat anti-murine mAb) for cytokine determinations were<br />

purchased from Pharmingen (San Diego, CA, USA). Atrial natriuretic factor (ANP) was<br />

delivered by Calbiochem (Schwalbach/Ts, Germany). CD11b-coated magnetic<br />

MicroBeads as well as FITC-labelled CD11b mAB were obtained from Miltenyi Biotec<br />

(Bergisch Gladbach, Germany). Other recombinant enzymes not further specified were<br />

purchased from Boehringer Mannheim (Mannheim, Germany) or Sigma (Deisenhofen,<br />

Germany).<br />

- 47 -


3.3 Cell culture materials<br />

Materials & Methods<br />

Cell culture plates (24 and 96 well), petri dishes and other plastic materials were<br />

purchased from Greiner (Frickenhausen, Germany). Cell culture medium RPMI 1640<br />

was purchased from BioWhittaker (Verviers, Belgium), DMEM HIGH GLUCOSE medium<br />

was from PAA laboratories (Pasching, Austria) and collagen was obtained from Serva<br />

(Heidelberg, Germany). Penicillin, streptomycin and FCS were bought from Gibco BRL<br />

Life Technologies (Eggenstein, Germany).<br />

3.4 Animals<br />

Specific pathogen-free male BALB/c, C57Bl/6 wild type mice, TNFR1 k.o. mice,<br />

TNFR2 k.o. mice and TNFR1R2 k.o. mice (approximately 25 g), originally provided by<br />

Dr. Horst Bluethmann, Basel, were from the in-house breeding stock at the University of<br />

<strong>Konstanz</strong>. LPR mice were from Harlan (Borchen, Germany). Inducible NFkB p65 k.o.<br />

mice (MXcrep65 k.o. and MXcrep65 wt) were a gift from Prof. Roland Schmid, München.<br />

Animals were held at 22°C and 55% humidity, given a constant day-night cycle of 12 h<br />

and fed a standard laboratory chow (Altromin 1310, Lage, Germany). All steps of animal<br />

handling were carried out according to the Guidelines of the National Institute of Health<br />

(NIH), the European Council (directive 86/609/EEC) and the national German authorities<br />

and followed the directives of the University of <strong>Konstanz</strong> Ethical Committee. Mice were<br />

starved overnight before the onset of experiments, which generally commenced at 8<br />

a.m.<br />

- 48 -


Methods<br />

3.5 Cell culture<br />

3.5.1 Cell lines<br />

HepG2 cells<br />

Materials & Methods<br />

HepG2 cells were maintained in RPMI 1640 medium containing 10% FCS. Cells<br />

were grown in 175 cm² flasks in humidified atmosphere at 37°C, 5% CO2, 21% O2 and<br />

74% N2. The day before the experiments were carried out, human hepatoma cells were<br />

harvested with trypsin/EDTA, centrifuged (200 × g, 4 min), resuspended in medium<br />

containing 10% FCS and plated in 24-well plates (10 5 cells/well).<br />

Hepa1-6 cells<br />

Hepa1-6 cells were maintained in DMEM HIGH GLUCOSE medium containing<br />

10% FCS. Cells were grown in 175 cm² flasks in humidified atmosphere at 37°C, 5%<br />

CO2, 21% O2 and 74% N2. The day before the experiments were carried out, murine<br />

hepatoma cells were harvested with trypsin/EDTA, centrifuged (200 × g, 4 min),<br />

resuspendet in medium containing 10% FCS and plated in 24-well plates (10 5 cells/well).<br />

AML-12 cells<br />

AML-12 cells were maintained in HAM’s F-12 medium containing 10% FCS. Cells<br />

were grown in 175 cm² flasks in humidified atmosphere at 37°C, 5% CO2, 21% O2 and<br />

74% N2. The day before the experiments were carried out, cells were harvested with<br />

trypsin/EDTA, centrifuged (200 × g, 4 min), resuspended in medium containing 10%<br />

FCS and plated in 24-well plates (10 5 cells/well).<br />

- 49 -


RAW cells<br />

Materials & Methods<br />

RAW cells were maintained in RPMI 1640 medium containing 10% FCS. Cells<br />

were grown in 175 cm² flasks in humidified atmosphere at 37°C, 5% CO2, 21% O2 and<br />

74% N2. The day before the experiments were carried out, cells were harvested with<br />

trypsin/EDTA, resuspended in medium containing 10% FCS and plated in 75 cm² flasks.<br />

3.5.2 Isolation and culture of primary cells<br />

Primary murine hepatocytes<br />

Isolation of hepatocytes from 8 weeks old mice was performed by the two-step<br />

collagenase perfusion method of Seglen (358) as modified by Klaunig (359, 360) and<br />

Leist (361). For some experiments, cells were additionally purified by centrifugation<br />

using a Percoll gradient modified from (362). To separate hepatocytes from remaining<br />

non-parenchymal cells, the pellet of the second centrifugation step (50 × g, 2.5 min) was<br />

resuspended in 16 ml Hanks´ Balanced Salt Solution (HBSS) and mixed with 32 ml of an<br />

isotonic Percoll solution (27.8 ml of 100% Percoll, 4.2 ml of 10-fold concentrated HBSS)<br />

followed by centrifugation at 800 × g for 5 min at room temperature. To remove<br />

remaining Percoll, the pellet was washed with HBSS by an additional centrifugation step<br />

(50 × g, 2.5 min). Hepatocytes were plated in 200 µl RPMI 1640 medium including 10%<br />

FCS in collagen-coated 24-well plates at a density of 8×10 4 cells per well. Cells were<br />

allowed to adhere to the plate for at least 4 hours before the medium was changed to<br />

RPMI 1640 without FCS. Incubation of murine hepatocytes with melphalan started 30<br />

min after medium exchange alone or in the presence of other mediators mentioned in<br />

the text. For some experiments Kupffer cells were depleted in vivo by intravenous<br />

injection of 150 µl liposome-encapsulated Clodronate one and two days before liver cell<br />

preparation. Clodronate-liposomes were prepared and injected as described earlier<br />

(363).Incubations were carried out at 37°C in an atmosphere of 40% O2, 5% CO2, 55%<br />

N2 and 100% humidity.<br />

- 50 -


Non-parenchymal liver cells<br />

Materials & Methods<br />

Non-parenchymal cells (NPCs) were plated in 100 µl RPMI 1640 / 10% FCS in 96<br />

well plates at a number of 10 5 per well. For standard experiments cells (mainly Kupffer<br />

cells) were allowed to adhere for 20 minutes before renewal of medium. Incubations<br />

were performed on the following day. For some experiments Kupffer cells were<br />

additionally purified as described by Doolittle et al. (364). After isolation cells were three<br />

times washed with PBS and incubated with CD11b-coated MicroBeads ® (10µl beads per<br />

10 7 total cells in a total volume of 100µl, 10 min, 4°C). After incubation, cells were<br />

washed by adding 900µl buffer with subsequent centrifugation at 300xg for 10 min, 4°C.<br />

After washing cells were resuspendet in buffer (10 8 total cells per 500 µl) and proceeded<br />

to magnetic separation (MS column type). Column was prepared according to manual<br />

instruction and cell suspension was allowed to pass magnetic field by gravity. Column<br />

was washed three times with 500 µl buffer before positive cells have been flushed out in<br />

a total volume of 1ml. After an additional washing step cells have been plated as usual<br />

in final concentration of 10 5 per well.<br />

For transwell experiments Kupffer cells were left to adhere for 24 hours on cell<br />

culture inserts with 1 µm pore size (Becton Dickinson, USA) and were placed in wells<br />

containing freshly isolated and adhered purified hepatocytes.<br />

3.6 Isolated liver perfusion<br />

3.6.1 Technical procedure of liver preperation and perfusion<br />

Upon a lethal intravenous injection with 150 mg/kg pentobarbital-natrium and 0.8<br />

mg/kg heparin, the vena portae and the vena cava inferior of the mouse liver were<br />

cannulated and ligated. After cannulation the organ was perfused blood-free before<br />

circulation has been closed, to guarantee a blood-free perfusion. Perfusion was carried<br />

out submarine (to avoid oedema formation) using a modified Krebs-Henseleit buffer [147<br />

mM NaCl; 5.36 mM KCl, 0.34 mM Na2HPO4; 0.44 mM KH2PO4; 0.77 mM MgSO4; 10<br />

mM D-Glucose; 9 mM HEPES; pH: 7.3; 325-330 mOsm] with a total volume of 25 ml<br />

buffer in a closed circulation mode under constant pressure conditions of 21.33 mm Hg.<br />

Osmolality of the perfusate fluid was set to 225-330 mOsm and checked before each<br />

individual experiment. The temperature of the perfusate was kept constant at 37°C,<br />

guaranteed by a separate thermo-circulation and oxygenation with pure oxygen at a<br />

pressure of 500 mbar was performed. During perfusion, the perfusate flow through the<br />

- 51 -


Materials & Methods<br />

liver as well as the pressure were constantly measured and recorded. Samples for<br />

metabolite and enzyme measurements were taken from the perfusate at different time<br />

points, as indicated in the text. All technical equipment and tubes for performing isolated<br />

mouse liver perfusion was delivered by Hugo Sachs Electronik (March-Hugstetten,<br />

Germany). For illustration see also figure 4.<br />

- 52 -


3.6.2 Treatment schedules<br />

Materials & Methods<br />

After lethal anaesthesia mice livers were immediately prepared and perfused with<br />

freshly oxygenated buffer to avoid ischemia. Total preparation time was below 5<br />

minutes, otherwise the experiment was cancelled. After closing the recirculation, organs<br />

were perfused for about 20 minutes to control leakage, perfusion parameters and quality<br />

of preparation. As soon as perfusate flow has stabilised, liver injury-inducing compounds<br />

were added and sampling of material started. Following injury-inducing models have<br />

been used:<br />

• Liver injury induced by anti-CD95: hepatic apoptosis mediated via CD95 was<br />

induced by application of agonistic anti-CD95 antibodies in concentrations of 100<br />

ng/ml given to the perfusate in a volume of 100 µl 0.1% HSA/saline at the<br />

beginning of experiment (t=0).<br />

• GalN/TNF-induced liver injury: GalN was given i.p. in doses of 500 mg/kg 30 min<br />

before lethal anaesthesia. TNF was given to the perfusate at a final concentration<br />

of 100 ng/ml in 100 µl 0.1% HSA/saline at the beginning of experiment (t=0).<br />

• GalN/LPS-induced liver injury: GalN was given i.p. in doses of 500 mg/kg 30 min<br />

before lethal anaesthesia. LPS was given in a final concentration of 1 µg/ml to the<br />

perfusate in a total volume of 300 µl sterile and pyrogen-free saline at the<br />

beginning of the experiment (t=0).<br />

• Melphalan-induced liver injury: Melphalan was prepared and added immediately<br />

to the perfusate in dose of 150 mg/kg at the beginning of the experiment (t=0)<br />

• Further compounds: carbohydrates and phorone were injected i.p. in doses<br />

indicated in the text/graph 30 min before lethal anaesthesia of animals. zVAD-fmk<br />

was added at concentrations indicated directly to the perfusate in a total volume<br />

of 100 µl 0.1% HSA/saline.<br />

• For depletion of Kupffer cells mice were injected two times three days and one<br />

day before the experiment with Chlodronate liposomes with a total volume of 300<br />

µl of provided liposome suspension according to Dr. Nico van Rooijen ((363) and<br />

personal communication) and Schümann et al. (365).<br />

- 53 -


3.6.3 Sampling of material<br />

Materials & Methods<br />

At the time points indicated (usually in 30 min interval) 100 µl of perfusate samples<br />

were taken, stored at 4°C and analysed immediately after the experiment (ALT and LDH<br />

measurement). In case of samples for TNF-measurement, samples were taken in<br />

shorter intervals as indicated and immediately frozen at -80°C. After the experiment<br />

livers were perfused for 10 s with cold perfusion buffer (PB, 50 mM phosphate buffer pH<br />

7.4, 120 mM NaCl, 10 mM EDTA), immediately excised and processed as follows:<br />

• Slices of the large anterior lobe were frozen in liquid nitrogen and stored at -80°C<br />

until the measurement of caspase-3,7-like activity.<br />

• For liver histology, liver specimen were immediately cut into 1 mm thick slices and<br />

fixed in phosphate-buffered neutral 4% formalin solution for light microscopy or<br />

frozen immediately on dried ice and stored at -80°C for cryoslices.<br />

3.7 Animal experiments<br />

3.7.1 Treatment schedules<br />

After treatment with liver injury-inducing compounds, animals were sacrificed by<br />

lethal anaesthesia to obtain samples at different times as described. Alternatively, the<br />

survival of animal was monitored over a period of at least three months.<br />

• Liver injury induced by anti-CD95: hepatic apoptosis mediated via CD95 was<br />

induced by application of agonistic anti-CD95 antibodies in doses of 2 µg/mouse<br />

given i.v. in a volume of 300 µl 0.1% HSA/saline.<br />

• GalN/TNF-induced liver injury: TNF was given i.v. in a dose of 2 µg/kg in 300 µl<br />

0.1% HSA/saline and the aminosugar GalN (500 mg/kg, given in 300 µl saline,<br />

i.p.) was administered 30 min before TNF to block hepatic transcription (8 hour<br />

model).<br />

• GalN/LPS-induced liver injury: after sensitisation with GalN as mentioned above,<br />

LPS was administered i.p. in a volume of 300 µl sterile saline in a dose of 1.5<br />

µg/kg (8 hour model).<br />

- 54 -


Materials & Methods<br />

• GalN/SEB-induced liver injury: after sensitisation with GalN as mentioned above,<br />

SEB was administered i.p. in a volume of 300 µl sterile saline in a dose of 2<br />

mg/kg (8 hour model).<br />

• Con A-induced liver injury: T cell-dependent liver injury was induced by Con A<br />

according to Tiegs et al. (366). Con A was injected i.v. into naive mice in a volume<br />

of 300 µl pyrogen-free saline at a dose of 25 mg/kg (8 hour model).<br />

• Further compounds: Melphalan was given in variable doses of 25, 50, 75 and 100<br />

µg/kg at different time-point before or after challenge, in a volume of 300 µl<br />

solvent.<br />

3.7.2 Sampling of material<br />

At the time-points indicated, mice were euthanized by i.v. injection of 150 mg/kg<br />

pentobarbital plus 0.8 mg/kg heparin and blood samples were obtained:<br />

• To assess the extent of liver damage, blood was withdrawn by cardiac puncture<br />

and subsequently centrifuged (5 min, 14,000 g, 4°C). ALT enzyme activity was<br />

measured in the plasma as described below.<br />

Blood samples for cytokine determinations were obtained either from the tail<br />

veins using heparinized syringes, or alternatively by cardiac puncture as<br />

described above, subsequently centrifuged (5 min, 14,000 x g, 4°C) and stored at<br />

-80°C.<br />

After blood withdrawal, livers were perfused for 10 s with cold perfusion buffer (PB, 50<br />

mM phosphate buffer pH 7.4, 120 mM NaCl, 10 mM EDTA), immediately excised and<br />

processed as follows:<br />

• Slices of the large anterior lobe were frozen in liquid nitrogen and stored at -80°C<br />

until the measurement of caspase-3-like activity.<br />

• For liver histology, liver specimen were immediately cut into 1 mm thick slices and<br />

fixed in phosphate-buffered neutral 4% formalin solution for light microscopy or<br />

frozen immediately on dried ice and stored at -80°C for cryoslices.<br />

- 55 -


3.8 Tumor model<br />

3.8.1 Tumor induction<br />

Materials & Methods<br />

12 to 15 days old C3H/HeN mice (Charles River, Sulzfeld, Germany) were injected<br />

intraperitoneally with 10µg/kg N-Nitrosodiethylamine- (Diethylnitrosamine; DEN) solution<br />

(Serva, Heidelberg, Germany). Animals were held at 22°C and 55% humidity under<br />

specific pathogen-free conditions, given a constant day-night cycle of 12 h and fed a<br />

standard laboratory chow (Altromin 1310, Lage, Germany). All steps of animal handling<br />

were carried out according to the Guidelines of the National Institute of Health (NIH), the<br />

European Council (directive 86/609/EEC) and the national German authorities and<br />

followed the directives of the University of Tübingen Ethical Committee and University of<br />

<strong>Konstanz</strong> Ethical Committee. Starting 20 weeks after tumor induction mice were killed by<br />

a lethal intravenous injection with 150 mg/kg pentobarbital-natrium and 0.8 mg/kg<br />

heparin and prepared for liver perfusion. Mice were starved overnight before the onset of<br />

experiments, which generally commenced at 8 a.m.<br />

3.8.2 Isolated liver perfusion<br />

Liver perfusion was carried out as described above. Prior to liver isolation tumor<br />

burden was documented by photographs. After cannulation of vena cava inferior and<br />

vena portae and subsequent perfusion with buffer to remove hepatic blood, a second<br />

picture was taken to document the perfusion of the tumors. Further perfusion procedure<br />

was carried out as indicated above. After the experiment, livers were excised out of<br />

mice, wet weight was measured and size of tumors and livers were documented by a<br />

picture taken on paper ruled in millimetre squares. Subsequently liver and tumor<br />

samples were taken and immediately frozen in liquid nitrogen for caspase-determination,<br />

frozen on dried ice for cryo-slices, or stored in freshly prepared 4% paraformaldehydesolution<br />

for histology.<br />

3.8.3 Treatment schedules<br />

After lethal anaesthesia mice livers were immediately prepared and perfused with<br />

freshly oxygenated buffer as soon as possible to avoid ischemia. After closing the<br />

recirculation, organs were perfused for about 20 minutes to control leakage, perfusion<br />

- 56 -


Materials & Methods<br />

parameters and quality of preparation. As soon as perfusate flow has stabilised, liver<br />

injury-inducing compounds were added and sampling of material started. Following<br />

treatment schedules have been investigated:<br />

• TNF: TNF was given to the perfusate at a final concentration of 100 ng/ml in 100<br />

µl 0.1% HSA/saline at the beginning of experiment (t=0).<br />

• Melphalan: Melphalan was prepared and added immediately to the perfusate in<br />

dose of 150 mg/kg at the beginning of the experiment (t=0)<br />

• Melphalan/TNF: Melphalan was prepared and added immediately to the perfusate<br />

in dose of 150 mg/kg at the beginning of the experiment (t=0), TNF was added to<br />

the perfusate at a final concentration of 100 ng/ml in 100 µl 0.1% HSA/saline 30<br />

min later<br />

• Control: 100 µl 0.1% HSA/saline was added at the beginning of the experiment<br />

(t=0)<br />

• Further compounds: carbohydrates were injected i.p. in doses indicated in the<br />

text/graph 30 min before lethal anaesthesia of animals.<br />

3.9 Cytokine determination<br />

Antibody pairs (specific rat anti-murine mAb) were purchased from Pharmingen<br />

(San Diego, CA, USA). Streptavidin-peroxidase was obtained from Jackson Immuno<br />

Research (West Grove, PA, USA) and the TMB liquid substrate system from Sigma<br />

(Deisenhofen, Germany). TNF and IL-10 were determined using a commercially<br />

available ELISA kit (OptEIA, Pharmingen, San Diego, CA. USA). The detection limits of<br />

the assays were 15 pg/ml for TNF and IL-10, 10 pg/ml for IL-4, IL-6 and IFN-γ.<br />

- 57 -


3.10 Light microscopy<br />

3.10.1 HE-staining<br />

Materials & Methods<br />

For light microscopy, liver samples were fixed in 4% buffered formalin and<br />

embedded in paraffin. Five-micrometer sections were cut (Biocut 2030, Reichert Jung,<br />

Germany) and stained with hematoxilin and eosin (Merck Biosciences, Darmstadt,<br />

Germany). Representative sections are shown.<br />

3.10.2 Tunnel-staining<br />

For light microscopy, liver samples were fixed in 4% buffered formalin and<br />

embedded in paraffin. Five-micrometer sections were cut (Biocut 2030, Reichert Jung,<br />

Germany) and stained with sheep-Digoxygenin-POD (Roche, Penzberg, Germany) for<br />

60 min at 37°C. After flushing with buffer, anti sheep-Peroxidase-conjugated antibody<br />

(Dianova, Hamburg, Germany) was added and slices were incubated for 60 min at 37°C.<br />

After flushing with buffer for 5 min, 3-amino-9ethyl-carbazole (Sigma, Deisenhofen,<br />

Germany) was added for 25 min before the slices were embedded in gelatine.<br />

Representative sections are shown.<br />

3.11 Measurement of enzyme activities<br />

3.11.1 Liver enzyme activities in plasma or samples<br />

The extent of liver damage was assessed by measuring plasma alanine<br />

aminotransferase (ALT) activity with an EPOS 5060 analyzer (Netheler & Hinz,<br />

Hamburg, Germany) according to the method of Bergmeyer (367-369) and calculated as<br />

U/l plasma.<br />

- 58 -


3.11.2 Lactate dehydrogenase activity<br />

Materials & Methods<br />

Lactate dehydrogenase (LDH) was determined in hepatocyte cell culture<br />

supernatants (S) and in the remaining cell monolayer (C) after lysis with 0.1% Triton X-<br />

100 according to Bergmeyer (367-369). The percentage of lactate dehydrogenase<br />

release was calculated from the ratio of S/(S+C). Alternatively LDH was measured using<br />

the cytotoxicity detection kit (Roche Diagnostics GmbH, Mannheim, Germany). The<br />

percentage of lactate dehydrogenase release was calculated equal to the calculation<br />

mentioned above.<br />

3.11.3 Caspase 3,7-like activity<br />

Cytosolic extracts from liver tissue were prepared by Dounce homogenization in<br />

hypotonic extraction buffer (25 mM HEPES, pH 7.5, 5 mM MgCl2, 1 mM EGTA, 1 mM<br />

Pefabloc® and pepstatin, leupeptin and aprotinin, 1 µg/ml each), subsequently<br />

centrifuged (15 min, 14,000 x g, 4°C) and stored at -80°C. To determine caspase-3-like<br />

protease activity, the fluorometric DEVD-afc cleavage assay was carried out according<br />

to the method originally described by Thornberry (370). Cytosolic extracts (10 µl,<br />

approximately 1 mg/ml protein as estimated with the Pierce-Assay (Pierce, IL, USA)<br />

were diluted 1:10 with substrate buffer (55 µM fluorogenic substrate DEVD-afc in 50 mM<br />

HEPES, pH 7.4, 1% sucrose, 0.1% CHAPS, 10 mM DTT). Generation of free 7-amino-4trifluoromethylcoumarin<br />

(afc) at 37°C was kinetically determined by fluorescence<br />

measurement (excitation: 385 nm; emission: 505 nm) using the fluorometer plate reader<br />

Victor2 (Wallac Instruments, Turku, Finland). The activity was calculated using serially<br />

diluted standards (0-5 µM afc). Control experiments confirmed that the activity was linear<br />

with time and with protein concentration under the conditions described above.<br />

Alternatively, to determine caspase-3-like and caspase 8 activity in vitro primary<br />

hepatocytes were lysed (freeze-thaw in lysis buffer: 25 mM HEPES pH 7.5, 5 mM<br />

MgCl2, 1 mM EGTA, 1 mM Pefablock and pepstatin, leupeptin and aprotinin, 1 µg/ml<br />

each, 0.1% Triton X-100), subsequent centrifuged (15 min, 14,000 g, 4°C), the<br />

supernatant diluted 1:10 in substrate buffer including DEVD-afc (caspase3) or IETD-afc<br />

(caspase-8) as substrate and generation of free afc determined using the fluorometer<br />

plate reader Victor 2 .<br />

- 59 -


Materials & Methods<br />

3.11.4 TNF-α-converting enzyme (TACE)-activity<br />

Recombinant TACE enzyme (Merck Biosciences, Darmstadt, Germany) was<br />

reconstituted in 1 ml of sterile water and aliquots were frozen at -20° C until usage as<br />

indicated in the data sheet. For photometric analysis, TACE was solved in a buffer<br />

containing 50 mM Tris-HCl, pH 7.4 50 mM NaCl and 4% glycerol. For monitoring TACE<br />

activity, an internally quenched fluorogenic substrate (Substrate IV, Merck Biosciences,<br />

Darmstadt, Germany) was used with an excitation maximum of ~320 nm and an<br />

emission maximum of ~420 nm (371). TACE inhibition experiments were carried out at a<br />

temperature of 37°C in a total volume of 1.0 ml containing 200 ng/ml of the recombinant<br />

protein, 5 µM of TACE substrate IV and various concentrations of a known inhibitor for<br />

TACE, TAPI-1 (Merck Biosciences, Darmstadt, Germany), or melphalan. Spectra were<br />

generated using a Luminescence Photometer LS 50B (Perkin Elmer, Rodgau-<br />

Jügesheim, Germany) with a slot of 5 nm.<br />

3.12 Determination of ATP<br />

Livers were perfused for 2 sec with ice-cold Somatic Cell ATP Releasing Reagent<br />

(Sigma, Germany) and homogenised by Dounce homogenisation. The 10% homogenate<br />

(in ATP Releasing Reagent) was centrifuged at 13,000 x g for 5 min at 4°C.<br />

Immediately, the supernatants were diluted and luminescence was measured in 96-well<br />

plates using an automated procedure (Victor² fluorometer plate reader (Perkin Elmer Life<br />

Sciences, Turku, Finland). Data were compared to calibration solutions and ATP<br />

contents are expressed in relation to protein content as percent of untreated control<br />

livers.<br />

For cellular ATP-determination of isolated hepatocytes, medium was replaced by<br />

150 µl ALP-lysis buffer (ATP Bioluminescence Kit HS II, Boehringer Mannheim,<br />

Germany) per well (24well format) and immediately frozen at -80°C. After thawing up,<br />

samples were diluted 1:10 in dilution buffer according to the standard protocol of the kit<br />

and subsequent measured in 96-well plates as mentioned above. Data were compared<br />

to calibration solutions and ATP contents are expressed in relation to protein content as<br />

percent of untreated control hepatocytes.<br />

- 60 -


3.13 Determination of glutathione<br />

Materials & Methods<br />

The amount of total glutathione (GSx = GSH + 2 x GSSG) in liver cells was<br />

quantified according to the enzymatic cycling method originally described by Tietze et al.<br />

(372) Isolated hepatocytes were incubated as described above. At different time-points<br />

indicated in the text, supernatant was removed and 100µl lysis-buffer containing 100µM<br />

HCl and 10mM EDTA were added and immediately frozen at -80°C. After thawing up,<br />

cellular material was separated from supernatant by centrifugation (5 min, 1,000 g, 4°C).<br />

150µl of the supernatant was removed and total glutathione was quantified with an ACP<br />

5040 analyser (Eppendorf, Hamburg, Germany).<br />

Alternatively the Glutathione Assay Kit from Cayman (Ann Arbor, MI, USA) was used.<br />

The assay was performed according to the manual instructions.<br />

3.14 Cytokine determination<br />

Antibody pairs (specific rat anti-murine mAb) were purchased from Pharmingen<br />

(San Diego, CA, USA). Streptavidin-peroxidase was obtained from Jackson Immuno<br />

Research (West Grove, PA, USA) and the TMB liquid substrate system from Sigma<br />

(Deisenhofen, Germany). TNF and IFN-γ were determined using a commercially<br />

available ELISA kit (OptEIA, Pharmingen, San Diego, CA. USA). The detection limits of<br />

the assays were 15 pg/ml for TNF and 10 pg/ml for IFN-γ.<br />

3.15 Cell-based ELISA<br />

This assay was performed as described previously (373). In order to assess the<br />

expression of membrane-bound TNF, Kupffer cells (10 5 cells/well) were left to adhere in<br />

96-well plates for 24 hours and were then either left untreated or were treated with LPS<br />

(10 ng/ml) or melphalan (200 µg/ml) for various times of incubation. Subsequently, cells<br />

were washed with Hanks Balanced Salt Solution (HBSS, BioWitthaker, Verviers,<br />

Belgium), fixed for 45 min at room temperature with HBSS + 1% paraformaldehyde,<br />

washed again, after which non-specific binding was blocked by means of incubating the<br />

wells with HBSS + 5% bovine serum albumin (Serva, Heidelberg, Germany).<br />

Subsequently, cells were incubated for 60 min at room temperature under mild shaking<br />

with either 30 µl/well of HBSS + 5% BSA, with 10 µg/ml of the 1F3F3 neutralising rat-<br />

- 61 -


Materials & Methods<br />

anti-mouse TNF mAb (374), with 10 µg/ml of the Sylvio neutralising sheep-anti-mouse<br />

TNF mAb, or with 10 µg/ml of an isotype-matched control rat IgM (Innogenetics, Ghent,<br />

Belgium), both diluted in HBSS + 5% BSA. After 2 washing steps, cells were incubated<br />

for 45 min under mild shaking with 3 mg/ml of a goat-anti-rat IgG-alkaline phosphatase<br />

conjugate (when first antibody 1F3F3) or rabbit-anti-sheep IgG-alkaline phosphatase<br />

conjugate (when first antibody Sylvio) and washed 2 times with HBSS + 5% BSA and 1<br />

time with 2.5 M diethanolamine, pH 9.5. Finally, the substrate solution consisting of 0.58<br />

mg/ml of the fluorescent substrate Attophos (Promega, Madison, WI, US) and of 2.4<br />

mg/ml of the endogenous phosphatase activity blocking agent levamisole (Sigma,<br />

Heidelberg, Germany) diluted in diethanolamine buffer was added. After 5 min,<br />

fluorescence was measured at excitation wavelength 485 nm and emission wavelength<br />

530 nm in the Victor² fluorometer plate reader (Perkin Elmer Life Sciences, Turku,<br />

Finland). Background values due to the unspecific control IgM antibody were subtracted<br />

from the anti-mTNF IgM settings.<br />

3.16 Western blot analysis<br />

Samples resulting from cellular extracts were loaded on a 12% polyacrylamide gel,<br />

separated under reducing conditions and subsequently blotted on a Hybond<br />

nitrocellulose membrane (Amersham-Buchler, Brauenschweig, Germany). In a Bio-Rad<br />

semi-dry blotter at 0.8 mA/cm² for 45 min. Homogenous transfer to the membrane was<br />

controlled by Ponceau red staining. Blots were blocked for at least 45 min with 5% nonfat<br />

dry milk in TBS/T containing 0.05% Tween-20 and subsequently incubated with antimouse<br />

TNF monoclonal antibody (Sylvio) (1:5,000 in TBS/Tween, 5% non-fat dry milk).<br />

After washing and incubation with a polyclonal IgG-horseradish peroxidase coupled<br />

secondary antibody (1:10,000 in TBS/Tween, 5% non-fat dry milk), the blots were<br />

developed by chemiluminescence method (ECL, Amersham-Buchler, Brauenschweig,<br />

Germany) following the manufacturer’s protocol.<br />

- 62 -


3.17 Statistics<br />

Materials & Methods<br />

All data are generally given as means ± SD or S.E.M. as indicated. Statistical<br />

differences were determined by one-way analysis of variance (ANOVA) followed by<br />

Dunnett multiple comparison test of the control vs. other groups. Statistical analysis that<br />

included all vs. all comparisons was done by the Tukey multiple comparison test. The<br />

statistic program InStat ® (GraphPad software, USA) was used for statistics and a p<br />

value


4 Results<br />

Results<br />

4.1 Characterisation of melphalan-induced cell death in primary murine<br />

liver cells<br />

Melphalan induced a concentration-dependent cytotoxicity in primary murine liver<br />

cells, with an EC50 of around 100 µg/ml (figure 5A), and lead to a steady increase of<br />

LDH release over control incubations 9 hours after drug exposure (figure 5B).<br />

The mode of cell death induced by this substance was identified as apoptosis, as<br />

characterised by typical nuclear alterations i.e. chromatin condensation and cellular<br />

alterations as rounding off, membrane blebbing and finally secondary lysis of apoptotic<br />

vesicles (figure 6A).<br />

- 64 -


Results<br />

- 65 -


Results<br />

Interestingly, at high melphalan concentrations a slightly different shape of cell death<br />

could be observed beginning 9 hours after melphalan exposure (figure 7).<br />

This shape of death was previously described in primary hepatocytes incubated<br />

with the plant lektin Con A and characterised as a kind of feathery structure with<br />

membrane obtrusions (361). This distinctive morphology could be observed till about 14<br />

hours following melphalan incubation before the cells ended up in apoptosis. In addition<br />

to the morphological findings an increase of caspase-3-like protease activity measured<br />

in liver cell lysates could be obtained. In lysates of cells incubated with 200 µg/ml<br />

melphalan the peak of caspase activity was observed between 12 and 18 hours (figure<br />

6B). Incubations with higher melphalan concentrations (300 and 400 µg/ml) revealed<br />

that these concentrations lead to a switch from apoptosis to necrosis as indicated by a<br />

different morphology and a lack of caspase-activity (data not shown). These results were<br />

also supported by the finding that the pan-caspase inhibitor zVAD was able to protect<br />

cells from melphalan-induced apoptosis at 200 µg/ml (figure 6C+D) but not at 400 µg/ml<br />

(data not shown). Due to these initial findings most of the experiments were performed<br />

with concentrations ranging from 50 to 200 µg/ml.<br />

- 66 -


Results<br />

4.2 The role of death receptors in melphalan-induced apoptosis<br />

Apoptosis is known be induced via an intrinsic (mitochondrial) and an extrinsic<br />

receptor-mediated pathway. Therefore, we have been investigated the effect of death<br />

receptors and their role in melphalan-mediated toxicity. Melphalan is known to induce<br />

interstrand and intrastrand crosslinks in the DNA which might be sufficient to induce<br />

apoptosis via an intrinsic apoptotic pathway. Since TNF, together with CD95 ligand,<br />

represents one of the most important known apoptosis-inducing cytokines in<br />

hepatocytes, we subsequently have been investigated the potential implication of the<br />

death receptors CD95 and TNFR1 in melphalan-induced apoptosis. To find out whether<br />

the death receptor CD95 has any influence on melphalan-mediated cell death in<br />

hepatocytes, we made use of mice expressing a non-functional CD95 (375). As such we<br />

did not detect any difference in sensitivity towards melphalan-induced apoptosis in cells<br />

derived from normal C57Bl/6 mice, as compared to cells derived from lpr-mice, lacking a<br />

functional CD95 receptor on the surface of immune cells and several other cell types<br />

including hepatocytes (figure 8). Furthermore, addition of an agonistic anti-CD95<br />

monoclonal antibody (Jo-2) had no enhanced effect on cell death when co-incubated<br />

with 200 µg/ml of melphalan (data not shown).<br />

- 67 -


Results<br />

TNF mediates its cellular effects mainly via two TNF receptors of which only<br />

TNFR1 is known to mediate apoptosis through an intracellular death domain whereas<br />

TNFR2 seems to fulfil other functions. Interestingly, it turned out that cells derived from<br />

TNFR1 knockout mice were fully protected from melphalan-induced cell death in<br />

concentrations up to 400 µg/ml (figure 9A). To verify if melphalan-induced toxicity is<br />

altered when TNFR2 is missing, cells were derived from TNFR2 knockout mice and<br />

incubated with melphalan. Unexpectedly, these cells were also protected from<br />

melphalan-induced toxicity (figure 9B), indicating a role for both TNF receptors. The<br />

observation that cells derived from TNFR1R2 double knockout mice are also protected<br />

is in line with previous observations (figure 9C).<br />

- 68 -


Results<br />

Last but not least, a third group of death receptors also belonging to the TNF<br />

receptor superfamily with its most prominent members TRAIL receptors 1 and 2 (DR4<br />

and DR5) has been shown to be able to induce death in hepatocytes. We therefore<br />

incubated cells derived from wild type mice with different concentrations of kiTRAIL<br />

alone or in combination with melphalan, but no additional effects could be observed by<br />

this agonist. Indeed, we were not able to induce death in primary liver cells by this ligand<br />

alone or in combination with transcriptional or translational inhibitors, like actinomycin D<br />

or cycloheximide. On the contrary, cell lines like Hep G2 cells were sensitive upon ligand<br />

binding with an EC50 of about 3 ng/ml after sensitisation with 100 µM cycloheximide<br />

(data not shown). These results contribute to recent publications demonstrating that<br />

DR4 and DR5 are not present on murine liver cells and stress the fact that the cellular<br />

effects observed after melphalan exposure are linked to a specific reaction involving<br />

TNF receptors. Therefore, we attempted to investigate the role of TNF in this setup.<br />

4.3 TNF mediates melphalan-induced toxicity<br />

Since we could demonstrate that both receptors for TNF play a crucial role in<br />

melphalan-mediated apoptosis, we subsequently have been investigated the effect of a<br />

neutralising monoclonal rat anti-mouse TNF antibody 1F3F3 on melphalan-induced<br />

apoptosis. This antibody has been demonstrated to neutralise both soluble and<br />

membrane bound TNF (374, 376). Neutralisation of TNF abrogated melphalanincreased<br />

LDH release completely to the level of control incubations (figure 9D)<br />

confirming the causal role of TNF and its receptors as demonstrated before.<br />

Complementary results could be obtained by using another polyclonal sheep-anti mouse<br />

TNF neutralising antibody (Sylvio, inhouse preparation). Assuming that TNF is the major<br />

mediator of melphalan toxicity, additional TNF should be able to further enhance the<br />

observed toxicity. But this was not the case. However, co-incubations of melphalan and<br />

soluble TNF did not enhance the observed toxicity compared to incubations with<br />

melphalan alone (figure 10).<br />

- 69 -


Results<br />

4.4 Kupffer cells are the source of TNF mediating melphalan-induced<br />

toxicity<br />

Since TNF seemed to be responsible for melphalan-induced apoptosis induction<br />

via interaction with TNF receptors and since no report has been shown that hepatocytes<br />

are able to produce TNF, which might lead to autocrine stimulation, the question<br />

remained what the source of TNF has been. Previous experiments showed that cells<br />

isolated according to the method of Seglen et al. are not a pure fraction of hepatocytes<br />

but are contaminated with up to 10% of non-parenchymal cells (NPCs), mainly Kupffer<br />

cells. These liver-specific macrophages have been shown to produce huge amounts of<br />

soluble TNF upon stimulation with e.g. endotoxin (figure 11) and therefore might also be<br />

able to produce sufficient amounts of TNF when contaminating primary hepatocyte<br />

cultures.<br />

- 70 -


Results<br />

To investigate this problem, we prepared highly purified primary hepatocytes<br />

upon depletion of Kupffer cells, achieved by intravenous injection of clodronate<br />

liposomes and additional selection of hepatocytes as described briefly in materials and<br />

methods. In incubations of these purified hepatocytes with melphalan no more toxicity<br />

was detected (figure 12A, white bars). However, hepatocyte apoptosis could be restored<br />

by the addition of exogenous TNF to melphalan (figure 12A, black bars). Moreover,<br />

hepatocyte apoptosis induced by 200 µg/ml of melphalan could also be restored by<br />

addition of increasing numbers of previously purified NPCs (mainly Kupffer cells) as<br />

shown in figure 12B. These results strongly indicate that this cell type is the source of<br />

TNF. To exclude that TNF production of Kupffer cells was due to LPS contamination of<br />

melphalan, we investigated LPS contamination in a LAL-test revealing that melphalan<br />

had only a weak contamination with LPS (11,5 ± 5,78 pg/ml). In addition we incubated<br />

melphalan with primary murine hepatocytes derived either from C3H/HeN-mice or<br />

C3H/HeJ-mice. The latter mouse strain is containing a mutation in tlr-4 leading to loss of<br />

function of this receptor. Lacking cytokine production and LPS-related signaling in<br />

response to endotoxin has been demonstrated for immune cells derived from these<br />

mice. Data obtained with liver cells of these mice are in line with the results from the<br />

LAL-test, showing no LPS effect influencing melphalan-induced toxicity (figure 12C).<br />

- 71 -


Results<br />

Notably, TNF as such failed to directly induce apoptosis in hepatocytes, unless<br />

transcriptional or translational inhibitors are also present. These findings thus indicate<br />

that in absence of Kupffer cells the simultaneous presence of melphalan (which is a<br />

transcriptional inhibitor) and TNF is a necessary condition to initiate hepatocyte<br />

apoptosis.<br />

4.5 Kupffer cells lead to enhanced TNF secretion after LPS but not after<br />

melphalan stimulation<br />

To investigate whether melphalan is able to induce TNF production in isolated<br />

Kupffer cells, NPCs of wild type mice were prepared and incubated with different<br />

concentrations of LPS or melphalan and secretion of soluble TNF was determined in the<br />

supernatants. As shown in figure 13A, secreted TNF was not detected in supernatants<br />

of Kupffer cells incubated with various concentrations of melphalan LPS (10 ng/ml)<br />

instead increased TNF production and secretion significantly. Interestingly, when coincubated<br />

with melphalan LPS-induced TNF production was totally abolished. Moreover,<br />

this LPS-induced TNF production could be inhibited by pre-treatment with<br />

- 72 -


Results<br />

pentoxyphylline, an unspecific inhibitor of phosphodiesterases (figure 13B), showing that<br />

the observed increase of soluble TNF release is related to LPS stimulation. Notably, this<br />

TNF release was also significantly reduced when LPS was co-incubated with melphalan<br />

on MH-S cells, a lung macrophage cell-line (ctrl:163 ± 2,3 pg/ml; LPS: 5,879 ± 574<br />

pg/ml; LPS/TAPI-1: 3,125 ±156 pg/ml; LPS/melphalan: 2,623 ± 125 pg/ml; melphalan:<br />

408,5 ± 48 pg/ml) (data not shown).<br />

- 73 -


Results<br />

To verify this result also in the whole organ we prepared livers from wild type mice<br />

and perfused them in presence of LPS or melphalan. TNF was determined in perfusate<br />

samples at different time points. As such, a weak induction of soluble TNF with a<br />

maximum between 90 and 120 min after addition of LPS was detected. When perfused<br />

with 150 mg/kg of melphalan only amounts of TNF below levels of untreated control<br />

livers could be detected (figure 13C). Moreover, mice injected with either LPS or<br />

melphalan or both displayed release of TNF in the circulation when challenged with<br />

melphalan or melphalan combined with LPS, whereas mice challenged with LPS alone<br />

displayed high amounts of serum TNF 90 min after LPS injection (figure 13D).<br />

TNF is known to be produced first as an active membrane-bound protein before<br />

being cleaved off via an also membrane associated matrix-metalloproteinase called<br />

TACE (TNF-alpha converting enzyme) and released as soluble pleiotropic factor. We<br />

investigated if the amount of TNF presentation was modified on the surface of Kupffer<br />

cells after melphalan stimulation. Therefore, Kupffer cells were prepared and incubated<br />

with either melphalan or LPS. Membrane-associated TNF was determined subsequently<br />

using a cell-based ELISA. In line with previous observations, treatment of purified<br />

Kupffer cells with 200 µg/ml melphalan led to a significantly increased expression or<br />

presentation of membrane-bound (figure 14B) but not of secreted TNF (below the<br />

detection limit of the murine TNF-ELISA of 20 pg/ml for all investigated time points).<br />

Such increased expression or presentation of membrane-bound TNF was measured<br />

from 30 min on after melphalan incubation (figure 14B). In contrast, LPS (10ng/ml)<br />

increased both the production of membrane-bound (figure 14A) and secreted TNF (data<br />

not shown) with clearly different kinetics from melphalan.<br />

- 74 -


Results<br />

Since no positive and negative controls are available for this cell-based ELISA, we<br />

performed Western blots of RAW-cells, a murine Kupffer cell line, stimulated with<br />

melphalan or LPS. In that set of experiments a maximum of membrane TNF (about 26<br />

kDa) was observed after 2 hours of incubation in lysates of RAW cells stimulated with<br />

25, 50 or 100 µg/ml of melphalan whereas untreated control cells did not (figure 15A).<br />

Less membrane TNF could be detected after 1 hour and 3 hours of incubation. When<br />

cells were stimulated with concentrations ranging from 12.5 µg/ml to 200 µg/ml,<br />

maximum expression of membrane TNF could be observed at 100 µg/ml (figure 15B).<br />

When cells were incubated with LPS (0.1 µg/ml; 1.0 µg/ml and 10 µg/ml) a band of 26<br />

kDa could also be detected which was not present in untreated control cells (figure 15B).<br />

- 75 -


Results<br />

4.6 Melphalan inhibits recombinant TACE activity in vitro<br />

Since melphalan has been shown to be able to abrogate LPS-induced TNF<br />

secretion in primary macrophages, while leading to significantly increased levels of<br />

membrane associated TNF on the surface of Kupffer cells, we were interested if<br />

melphalan might be able to exert any inhibitory effect on TACE enzyme activity.<br />

Therefore, we set up a fluorimetric enzyme activity test for TACE using the recombinant<br />

protein and a specific chromophore-coupled substrate as described in the materials &<br />

methods section. To test the quality of the enzymatic assay, we measured TACE-activity<br />

in the presence of a known inhibitor for TACE, TAPI-1, with an IC50 of about 10 nM. As<br />

such, we found an IC50 of about 20 nM in our enzymatic assay for this inhibitor in three<br />

independent sets of experiments (figure 16A), demonstrating the quality of this assay.<br />

Incubations done in the presence of various concentrations of melphalan revealed that<br />

the study drug was indeed able to inhibit TACE-activity in a concentration range that fit<br />

in very well to results obtained in vitro before (figure 16B), stressing the hypothesis that<br />

TACE was also inhibited in our cellular assays.<br />

- 76 -


Results<br />

4.7 Melphalan is able to block TNF-dependent liver injury in vivo<br />

To investigate the role of melphalan-mediated inhibition of TACE enzyme activity in<br />

the living animal, we performed in vivo experiments, using known mouse models of<br />

cytokine-mediated liver injury. If melphalan is also able to block TACE activity and the<br />

release of soluble TNF in vivo, then these mice should at least be partially protected<br />

from cytokine-mediated liver injury.<br />

4.7.1 GalN/TNF and GalN/LPS model<br />

TNF has been shown to be hepatotoxic in man and mice (377). Mice can be<br />

dramatically sensitised towards TNF by viral or bacterial infection or chemicals such as<br />

galactosamine (GalN), resulting in an up to 10 4 -fold increased lethality (378) and<br />

hepatotoxicity (379). The amino sugar GalN selectively depletes uridine in hepatocytes<br />

(380) and therefore causes a selective transcriptional inhibition in the murine liver. The<br />

GalN/TNF model uses exactly the liver-specific inhibitory capacity of GalN to induce a<br />

liver specific toxicity in mice by activating apoptosis in the liver through exogenously<br />

applicated TNF. The GalN/LPS model is a more complex scenario for GalN/TNF<br />

mediated liver failure. The liver harbours the largest pool of resident macrophages in the<br />

body, i.e. Kupffer cells and thus it is a potent TNF-producing organ (381). Kupffer cells<br />

within the liver may produce TNF after direct stimulation, e.g. with LPS or after<br />

interaction with activated lymphocytes in a more complicated immunological setting<br />

(382) leading to a pathological situation, which finally leads to liver failure in GalNsensitised<br />

mice. Passive immunisation of mice against TNF prevented apoptosis (383)<br />

as well as the release of transaminases being a late hallmark of liver damage (383,<br />

384).<br />

Mice challenged with GalN (500 mg/kg) and LPS (Salmonella abortus equi 1.5<br />

µg/kg) displayed significantly enhanced liver-toxicity over controls (serum ALT-activity:<br />

19.6 ± 10 U/L; data not shown), eight hours after challenge. Toxicity was characterised<br />

by a significantly enhanced ALT activity in serum, as compared to controls obtained<br />

eight hours after challenge (figure 17A). In addition, mice displayed a significantly<br />

enhanced TNF release 90 min after challenge. Animals pre-treated with 100 mg/kg<br />

melphalan instead displayed significantly decreased levels of ALT in serum, as<br />

compared to mice challenged with GalN/LPS alone. Also TNF levels were significantly<br />

reduced in these pre-treated mice (figure 17B).<br />

- 77 -


Results<br />

Interestingly, this effect could also be observed when GalN/LPS-challenged mice<br />

were pre-treated with 75 mg/kg or 50 mg/kg melphalan (figure 17C+D). This inhibitory<br />

effect of melphalan could also be observed when GalN/LPS-challenged mice were<br />

injected with melphalan (50 mg/kg – 100 mg/kg) at the same time as GalN/LPS or up to<br />

30 min after GalN/LPS-challenge (figure 17E+F). When melphalan was given one hour<br />

after GalN/LPS-challenge, ALT release was still reduced, but TNF release was shown to<br />

be enhanced (figure 17E+F). This finding is of further interest, because we proposed<br />

- 78 -


Results<br />

that the protective effect of melphalan is only due to its ability to abrogate TNF-release.<br />

But in this case soluble TNF was present in the circulation while ALT-release was still<br />

reduced.<br />

To investigate this effect, we injected mice with GalN (500 mg/kg) /TNF (10 µg/kg)<br />

with and without pre-treatment of melphalan (100 mg/kg). Interestingly, these mice<br />

displayed also a significantly decreased plasma level of ALT activity 8 hours after<br />

challenge (figure 18A). Since this result was not expected we were further interested if<br />

other parameters in this well-characterised model are also altered. Therefore we<br />

determined caspase-activity in liver samples collected after the experiment and found<br />

significantly reduced levels of downstream caspases in melphalan pre-treated mice<br />

(figure 18B). Finally, we determined plasma activity of IFNγ eight hours after challenge,<br />

but no significant difference between GalN/TNF-treated mice and mice additionally pretreated<br />

with 100 µg/mg of melphalan was found (figure 18C).<br />

- 79 -


4.7.2 GalN/SEB<br />

Results<br />

The superantigen SEB exerts its toxicity through activation of a subset of T cells<br />

and subsequent production of TNF (385-389). Essential for the induction of apoptotic<br />

liver damage by SEB is the presence of transcriptional (386) or translational (365, 390)<br />

inhibitors. Dendritic cells are considered essential for clonal expansion of Vß8-specific T<br />

cells (391) and local expression of cytokine mRNA within the spleen (392).<br />

Macrophages, especially Kupffer cells, have also been shown to contribute essentially to<br />

this scenario being important for TNF production (365).<br />

Animals injected with GalN (700 mg/kg) and SEB (2 mg/kg) showed a severe liver<br />

toxicity eight hours after challenge, as shown by serum ALT activity (figure 19A). In<br />

addition, TNF could be detected in serum collected 90 min after challenge (figure 19B)<br />

and enhanced levels of IFN-γ could be detected eight hours after challenge (figure 19C).<br />

In contrast to these results, animals pre-treated with 100 mg/kg of melphalan one hour<br />

before SEB injection displayed significantly reduced levels of ALT (figure 19A), TNF<br />

(figure 19B) and IFN-γ (figure 19C). However, the observed spleen hypertrophy could<br />

not be blocked by melphalan (data not shown).<br />

- 80 -


4.7.3 The Con A model.<br />

Results<br />

Concanavalin A (Con A) is a plant lectin that activates T cells in vitro and in vivo.<br />

When injected into mice, Con A triggers a selective liver injury (366) depending on the<br />

release of the cytokines TNF (393) starting 90 min after challenge, IFN-γ (382) and IL-4<br />

(394). Further, the Con A model involves an early damage of sinusoidal endothelial cells<br />

(298, 395) and requires a cross talk between macrophages and T cells (365, 382). In the<br />

Con A model, both necrotic and apoptotic hepatocyte demise with or without a<br />

contribution of caspases have been described (297, 366, 393, 396, 397). Interestingly,<br />

hepatotoxicity induced by Con A does not require any kind of additional sensitisation and<br />

involves both TNF receptors, TNFR1 and TNFR2, as shown by studies using tnfr1 or<br />

tnfr2 gene deficient mice. Moreover, in contrast to all other TNF-dependent models of<br />

liver injury requiring sensitisation by transcriptional inhibitors, Con A-induced liver injury<br />

is independent of caspase-3-like protease activity, as shown by studies using the pancaspase<br />

inhibitor benzoyloxycarbonyl-val-ala-asp-fluoromethylketone (zVADfmk) (297).<br />

Animals injected with 25 mg/kg of ConA displayed fulminant liver toxicity eight<br />

hours after the treatment, as assessed by ALT-release in the serum (figure 20A). TNF<br />

release (90 min after challenge) (figure 20B) and IFNγ-release (8 hours after challenge)<br />

(figure 20C) in the circulation could be also detected in these mice. Melphalan pretreatment<br />

(100 mg/kg) instead, totally abrogated release of cytokines (TNF and IFN-γ)<br />

(figure 20B+C) and hepatotoxicity was significantly reduced (figure 20A). Since this<br />

model involves a cross talk between macrophages and T cells, we were interested if the<br />

well-characterised cytokine profile was altered upon melphalan pre-treatment.<br />

Therefore, we injected mice with 25 mg/kg of ConA with and without pre-treatment of<br />

100 mg/kg of melphalan for 1 hour and determined IL-4, IL-10, IL-1ß, IL-2 and IFNγ in<br />

blood samples collected after 90 min, 2 hours, 3 hours and 8 hours, respectively. As<br />

expected melphalan pre-treatment altered the whole cytokine profile of this model with<br />

TNF, IL-1ß and IFNγ being significantly reduced and IL-10 being significantly enhanced<br />

as compared to Con A treatment alone.<br />

- 81 -


Results<br />

- 82 -


4.7.4 The CD95 model<br />

Results<br />

This model has been used because liver failure induced by injection of agonistic<br />

antiCD95 antibody has been shown to be totally independent of TNF and its<br />

corresponding TNF receptors. The application of the activating anti-CD95 antibody Jo-2<br />

in naive mice leads to lethal liver destruction within hours, due to massive caspasemediated<br />

apoptosis of hepatocytes (398, 399). Importantly, this effect does not require a<br />

sensitisation of the animals and is seen when instead of anti CD95, the CD95L is used<br />

or endogenously generated (400, 401). Other organs than the liver have also been<br />

reported to be affected by anti CD95 (402, 403), but this depends largely on the<br />

antibodies and mouse strains used and therfore is of minor importance when Jo-2 was<br />

used (404). CD95 is also highly expressed in sinusoidal endothelial cells (405) and a<br />

prominent damage of these cells precedes the onset of liver damage in this model (406).<br />

Mice injected with agonistic αCD95 antibody (2 µg/mouse) displayed a significantly<br />

enhanced hepatotoxicity eight hours after challenge (figure 21A). As expected,<br />

melphalan pre-treatment (100 mg/kg, 60 min) was not able to block CD-95 mediated cell<br />

death, but was able to reduce serum IFN-γ significantly detected 8 hours after αCD-95<br />

challenge (figure 21B).<br />

- 83 -


Results<br />

4.7.5 Melphalan and melphalan/TNF model<br />

Since melphalan has been found to induce apoptosis in primary liver cells and in<br />

isolated perfused organs, we were interested, if melphalan is also able to induce liver<br />

toxicity in vivo. Therefore, we injected mice i.p. with various doses of melphalan, ranging<br />

from 25 mg/kg up to 150 mg/kg. Unfortunately, we were not able to detect any liver<br />

toxicity eight hours after challenge or in animals died earlier (data not shown). Mice<br />

injected with high doses of melphalan displayed renal alterations associated with bloody<br />

urine and black coloured bladders, intestinal toxicity associated with midgut changes<br />

and diarrhoea. Lowered body temperature as compared to untreated control animals<br />

and impaired balance was also observed. Therefore, we investigated if mice challenged<br />

with lower doses of melphalan (25 mg/kg – 75 mg/kg) and TNF (10 µg/kg) display any<br />

hepatic toxicity. But these experiments also failed (data not shown). Since no extended<br />

knowledge could have been expected from further in vivo experiments compared to data<br />

obtained in vitro no further investigations have been undertaken, also with respect to<br />

ethical guidelines for animal experiments.<br />

4.8 Cell-cell contact is necessary between Kupffer cells and hepatocytes<br />

Previous results indicated that melphalan was able to induce enhanced<br />

presentation of TNF on macrophages and inhibited the cleavage off the membraneanchor<br />

through TACE. On the other hand Kupffer cells or at least the presentation of<br />

TNF on the surface of Kupffer cells was shown to be essential for apoptosis induction by<br />

melphalan. So we were further interested if membrane-bound TNF indeed was<br />

responsible for mediating melphalan toxicity or if at least very small amounts of soluble<br />

TNF, undetecable in ELISA, were sufficient to stimulate TNF receptors in order to induce<br />

apoptosis in hepatocytes. Therefore, we performed transwell experiments, in which the<br />

Kupffer cells were cultivated in cell culture inserts that were spatially separated from the<br />

hepatocytes cultivated in the bottom wells. As shown in table 4, melphalan alone was<br />

not sufficient to induced cytotoxicity in hepatocytes when Kupffer cells were present but<br />

separated in cell culture inserts. On the other hand, cytotoxicity occurred when cell-cell<br />

contact between both cell types was enabled indicating that this contact between both<br />

cell types was crucial for apoptosis induction.<br />

- 84 -


Results<br />

4.9 Role of TNF receptors on Kupffer cells<br />

Since both TNF receptors have been shown to play a crucial role in mediating<br />

melphalan-induced apoptosis of hepatocytes, and Kupffer cells have been demonstrated<br />

to be necessary for melphalan-induced TNF-mediated cell death, the question was<br />

raised, whether both receptor types are necessary on both cell types. Since TNFR2 is<br />

lacking an intracellular death domain and therefore does not play a decisive role in<br />

receptor-mediated caspase activation, we have been interested if this receptor anyhow<br />

is only important in Kupffer cell signaling. TNFR1 instead, containing an intracellular<br />

death domain, would have been sufficient for transmitting the death signal given by TNFpresenting<br />

macrophages in this scenario. This alternative hypothesis would have also<br />

included a role for both TNF receptors, but on separate cell types; TNFR1 on the<br />

hepatocyte and TNFR2 on the Kupffer cell.<br />

- 85 -


Results<br />

To investigate this theory we prepared hepatocytes and Kupffer cells from different<br />

mouse strains, purified them and performed co-culture experiments in all possible kinds<br />

of variations using wild type, TNFR1, TNFR2 and TNFR1R2 knockout mice. The results<br />

obtained in these experiments are summarized in figure 22. In detail, hepatocytes<br />

isolated from wild type mice and co-cultured with Kupffer cells derived either from<br />

TNFR1 or TNFR2 or TNFR1R1 knockout mice died in response to melphalan,<br />

independent of the expression of TNF receptors. On the contrary, hepatocytes derived<br />

from TNFR1R2 knockout mice died not when co-cultured with different Kupffer cells and<br />

melphalan (figure 22 B).<br />

In conclusion these results demonstrated that both TNF receptors contribute to the<br />

toxicity on the surface of hepatocytes, while the expression of TNF receptors on the<br />

surface of Kupffer cells did not play a decisive role in mediating melphalan induced<br />

toxicity.<br />

- 86 -


Results<br />

4.10 Melphalan-induced toxicity on Kupffer cells<br />

Since we have been demonstrated that Kupffer cells play a crucial role in<br />

mediating melphalan-induced toxicity via both TNF receptors, the question remained<br />

how purified Kupffer cells respond to various melphalan concentrations.<br />

To investigate this problem, we isolated primary murine liver cells and purified<br />

Kupffer cells as described above. Cells were incubated with different concentrations of<br />

melphalan, varying from 6.25 µg/ml to 400 µg/ml of melphalan and cell viability was<br />

determined by Alamar blue assay, as well as by LDH release using the cytotoxicity<br />

detection kit (Roche diagnostics). As we could demonstrate, isolated Kupffer cells died<br />

in response to melphalan with a sensitivity comparable to hepatocytes (EC50 = 130<br />

µg/ml) (figure 23A). Interestingly, Kupffer cells isolated from mice deficient for both TNF<br />

receptors, displayed a similar concentration response curve, with an EC50 of 75 µg/ml<br />

(figure 23B), indicating that this cell death was independent of TNF receptor<br />

engagement. In kinetic experiments with Kupffer cells derived from wild type mice we<br />

could demonstrate that cells have been intact in the first nine hours following melphalan<br />

treatment since no LDH could be detected in the supernatant before that time point (data<br />

not shown). Having a closer look on the mode of cell death we incubated cells from wild<br />

type mice with melphalan in the presence and absence of 1 µM zVAD. Interestingly, no<br />

beneficial effect of zVAD co-incubation was observed (figure 23C), and corresponding to<br />

this no caspase-3-like protease activity could be determined (data not shown).<br />

- 87 -


Results<br />

These data stress the fact that Kupffer cell toxicity was independent of caspases.<br />

Nevertheless, since we have been measured onset of caspase activity already at about<br />

nine hours in primary liver cell cultures with a caspase-specific activity of about 120<br />

pmol/mgxmin after 12 hours, this Kupffer cell death was not decisive for the toxic effects<br />

on hepatocytes observed in primary liver cell culture. In addition, in presence of anti-TNF<br />

antibodies Kupffer cell death also occured, but hepatocytes were protected from induced<br />

toxicity in primary liver cell culture (data not shown).<br />

4.11 Melphalan-induced hepatotoxicity in the presence and absence of<br />

Kupffer cells<br />

In previous experiments we have demonstrated that the presence of Kupffer cells<br />

is a necessary condition to induce hepatocyte apoptosis by melphalan. Under Kupffer<br />

cell depleted conditions no hepatocyte death could be observed, but toxicity could be<br />

restored by adding small amounts of soluble TNF (figure 12A). This observation raised a<br />

number of questions concerning the kinetics of apoptosis induction, the kinetics of<br />

caspase activation and the crucial involvement of TNF receptors. To address these<br />

questions we prepared hepatocytes from Kupffer cell-depleted mice, additionally purified<br />

them as described in materials and methods and incubated them as indicated.<br />

Interestingly, it turned out that apoptosis induced by melphalan and soluble TNF<br />

under Kupffer cell depleted conditions showed clearly different kinetics, as compared to<br />

apoptosis induced in the presence of Kupffer cells with melphalan alone (figure 24A+B).<br />

Also the onset and the maximum of caspase-3-like activity differed in both settings<br />

(figure 24C). In the presence of Kupffer cells and the absence of exogenous TNF a<br />

maximum of caspase-3-like activity was found 18 hours after melphalan stimulation with<br />

a specific activity of about 200 pg/mlxmin, whereas in the absence of Kupffer cells and<br />

the presence of exogenous TNF the highest caspases-3-like protease activity was<br />

detected seven hours after melphalan/TNF stimulation with an activity of about 300<br />

pg/mlxmin (figure 24C). LDH release was found to be comparable in both settings<br />

showing a maximum between 18 and 24 hours following melphalan or melphalan/TNF<br />

stimulation (data not shown). The amount of cell death assessed was also comparable<br />

in both settings (about 80%) (figure 24A+B). Interestingly, the calculation of effective<br />

melphalan concentration revealed that in the presence of Kupffer cells and the absence<br />

of exogenous TNF EC50 was nearly doubled, as compared to the setting under Kupffer<br />

cell-depleted conditions (figure 24A+B). In addition, also the IC50 values for inhibition of<br />

cell-death by the pan-caspase inhibitor zVAD were completely different, ranging from 6<br />

- 88 -


Results<br />

nM in the situation of Kupffer cell absence to 100 nM in the situation of Kupffer cell<br />

presence (figure 24D).<br />

To address the question how TNF receptors were involved in melphalan/TNFinduced<br />

cell death, hepatocytes from different mouse strains depleted of Kupffer cells<br />

were prepared. As shown in figure 24E, cells lacking TNFR1 were still protected from<br />

melphalan/TNF-induced cell death, while cells derived from tnfr2 gene deficient mice<br />

were susceptible to melphalan/TNF-induced cell death (figure 24F). Notably, cells<br />

lacking both TNF receptors were also protected from apoptosis in this setting (data not<br />

shown).<br />

- 89 -


Results<br />

It is also worth noting that cells derived from macrophage-depleted wild type or<br />

TNFR2 gene deficient mice were also susceptible to apoptosis induction by melphalan<br />

and human TNF which is known to interact only with murine TNFR1 and not with murine<br />

TNFR2 (data not shown). These results were in line with previous findings, but again<br />

raised the question how TNFR2 was involved in this scenario when Kupffer cells were<br />

present.<br />

4.12 Melphalan induces TNF-mediated hepatotoxicity in situ<br />

In previous experiments we could demonstrate that melphalan-induced hepatocyte<br />

death in vitro was mediated by membrane TNF and both TNF receptors in the presence<br />

of Kupffer cells. It could also be demonstrated that due to the inhibition of TACE activity<br />

by melphalan no soluble TNF was released or at least in amounts being below the<br />

detection limit of the ELISA. In animal experiments melphalan pre-treatment prevented<br />

TNF-mediated liver failure in mice, but melphalan alone or in combination with<br />

exogenous TNF did not induce enhanced liver apoptosis or necrosis.<br />

In order to investigate if melphalan induced TNF-mediated hepatocyte apoptosis<br />

was only occurring under in vitro conditions, we decided to confirm the results obtained<br />

in vitro also under conditions of isolated liver perfusion. Isolated liver perfusions with<br />

melphalan and TNF represent a model being closely related to the situation in patients<br />

undergoing IHP for tumor therapy.<br />

In this model, control organs did not undergo a significant hepatotoxicity for up to<br />

540 min as indicated by the absence of ALT (figure 25A ▲) and LDH release (data not<br />

shown) in the perfusate. In addition, no caspase 3-like-activity was measured in liver<br />

samples collected after the experiment (data not shown) and liver architecture was<br />

found to be normal as assessed in H.E. staining and TUNNEL staining of liver slices<br />

collected from the perfused organs. In contrast, perfusion with 150 mg/kg of melphalan<br />

induced a significant hepatotoxicity, as evidenced by the release of ALT in organs<br />

derived from wild type (figure 25A ■) but not from knockout mice lacking TNFR1 (figure<br />

25B ×), TNFR2 (figure 25B ○), or both TNF receptors (data not shown). Moreover, the<br />

causal role of TNF action and crucial role of Kupffer cells was confirmed in livers derived<br />

from Kupffer cell depleted wild type (figure 25C ■), TNFR1, TNFR2 and TNFR1R2<br />

knockout mice (data not shown). In none of these settings melphalan induced significant<br />

liver failure as indicated by lacking release of ALT in the perfusate and the absence<br />

caspase-3-like activity in liver homogenates of perfused livers (figure 25 D).<br />

- 90 -


Results<br />

In line with these biochemical observations liver architecture was found to be<br />

normal up to 8 hours with some light necrotic areas occasionally seen in the periportal<br />

fields, but not in the central vein areas (figure 26). Analogous to these findings, organs<br />

isolated from wild type mice displayed a significant disturbance of liver architecture after<br />

360 min of perfusion with melphalan (figure 26), which was characterised by endothelial<br />

damage, oedema formation and apoptotic chromatin condensation in a subset of<br />

hepatocytes.<br />

- 91 -


Results<br />

- 92 -


Results<br />

Isolated liver perfusion with melphalan and TNF in the absence of Kupffer cells<br />

was also sufficient to reproduce results obtained in vitro. The onset of toxicity as well as<br />

the course of perfusion and the maximum of ALT release in the perfusate was<br />

comparable in both settings, perfusion with melphalan in the presence of Kupffer cells<br />

and perfusion with melphalan and exogenous TNF in the absence of Kupffer cells (figure<br />

25 C). In case of Kupffer cell depletion a maximum of about 2,800 U/L was reached after<br />

6 hours of perfusion, whereas in livers containing Kupffer cells a maximum of ALTrelease<br />

was observed after 6.5 hours with a total amount of about 1,500 U/L (figure 25<br />

C). Additionally, we could also demonstrate that livers derived from TNFR1 (figure 25 C<br />

♦) or TNFR1R2 knockout mice (data not shown) were fully protected up to 9 hours of<br />

perfusion. On the contrary, livers derived from TNFR2 knockout mice perfused with<br />

melphalan and TNF displayed an enhanced ALT release in the perfusate starting from<br />

330 min on and ending up with a total release of about 2,700 U/L after 420 min, as<br />

compared to the release of less than 100 U/L by untreated control livers,<br />

melphalan/TNF-treated TNFR1-deficient livers (figure 25 C▼) or TNF-treated TNFR2deficient<br />

livers (data not shown). In addition to these results, caspase-3-like activity<br />

measured in liver homogenates of tissue samples collected after the individual<br />

experiments was only found in livers perfused with melphalan and exogenous TNF, not<br />

in livers perfused with melphalan alone. Also no caspase activity could be obtained in<br />

livers derived from mice lacking TNFR1 either perfused with melphalan alone or with<br />

melphalan and TNF together (figure 25 D).<br />

These experiments demonstrated that the in vitro results were reproducible in the<br />

whole organ, a model being closely related to the clinical situation of the local<br />

therapeutic liver perfusion with melphalan. The crucial role of TNFR2 in melphalan<br />

induced liver toxicity could be also demonstrated in the whole organ.<br />

Since very less models of liver injury or pathologic situations have been described<br />

demonstrating a crucial role for TNFR2 in liver apoptosis, we were further interested how<br />

TNFR2 might contribute to this scenario in melphalan induced hepatocyte apoptosis in<br />

vitro.<br />

- 93 -


Results<br />

4.13 Modulation of intracellular signaling by various inhibitors<br />

As has been shown in the initial section of this chapter, melphalan/TNF-induced<br />

toxicity was strongly dependent on the activation of caspases, mainly driven by the<br />

activation of TNFR1. Since it is known that apoptosis via TNFR1, in contrary to CD95,<br />

has been only induced in the presence of transcriptional or translational inhibitors,<br />

melphalan seemed to be able to act in a bifunctional manner, blocking cellular<br />

transcription and/or translation on the one hand and triggering TNF-presentation on the<br />

other hand, which in turn lead to the activation of hepatocyte apoptosis. It is also known<br />

that TNF receptors mediate a variety of cellular stimuli upon binding to their specific<br />

ligand, including several biochemical control centres like NFκB, p38 MAPK or JNK.<br />

Because these molecules have been shown to be involved in many cellular processes, a<br />

number of inhibitors have been generated and made commercially available for selective<br />

inhibition of cellular pathways, in order to prove their crucial role in various cellular<br />

processes.<br />

4.13.1 Role of NFκB<br />

One of the most interesting molecules that has been shown to be activated upon<br />

TNF stimulation is NFκB. In order to investigate the role of NFκB, we used an<br />

ammonium salt pyrrolidinedithiocarbamate (PDTC) which has been demonstrated to<br />

prevent NFκB translocation as well as induction of nitric oxide synthase by inhibiting<br />

translocation of NOS mRNA (407). To test the influence of PDTC, we incubated primary<br />

murine liver cells with concentrations from 0.1 µM up to 100 µM of PDTC alone or in<br />

combination with melphalan in concentrations from 12.5 µg/ml to 400 µg/ml. But as<br />

shown in figure 27A (data shown for 10 µM) PDTC had at least no inhibitory influence on<br />

melphalan-induced toxicity. On the contrary, inhibition of NFκB enhanced melphalan<br />

toxicity at low concentrations.<br />

- 94 -


Results<br />

In order to verify these results in a separate experimental system, we made use of<br />

a special transgenic mouse strain, an inducible NFκB p65 (MXcrep65) knockout mouse.<br />

These mice have been shown to loose NFκB p65 on day 4 till day 6 after intravenous<br />

application of polyIC (250 µg/mouse) (Nektarios Dikopoulos, personal communication).<br />

Interestingly, we found out that MXcrep65 -/- mice or liver cells and isolated perfused livers<br />

derived from those mice were sensitive to TNF alone without pre-incubation of<br />

transcriptional inhibitors such as ActD or GalN when treated on day 4 or day 5 following<br />

polyIC injection. This could be shown by toxicity assays and caspase-3-like protease<br />

activity (data not shown). On the contrary, when mice or liver cells derived from these<br />

mice were not injected with polyIC and stimulated with TNF no apoptosis occured. Also<br />

no apoptosis was observed when mice were injected with polyIC and cells or organs<br />

- 95 -


Results<br />

were treated on day 1 or day 2 after the injection. MXcrep65 +/+ mice, carrying a nonfunctional<br />

cre-lox-system, or cells and organs derived from those mice were not<br />

sensitive to TNF alone only in combination with transcriptional or translational inhibitors<br />

(data not shown).<br />

Primary liver cells derived from MXcrep65 -/- mice and incubated with melphalan as<br />

usual, showed no significant difference in apoptosis induction as compared to cells<br />

derived from MXcrep65 +/+ mice or normal C57Bl/6 mice (figure 27B). Unfortunately, due<br />

to a limited number of mice we got for a co-laborational project we were not able to<br />

reproduce these effects in isolated perfused mouse livers Nevertheless, the results<br />

obtained in primary cell culture are in line with our working hypothesis. To investigate if<br />

NFκB translocates to the nucleus upon melphalan stimulation, we performed NFκBbandshift<br />

assays in primary murine liver cells derived from normal C57Bl/6 mice. Data<br />

obtained by this assay indicated that NFκB p65 shows a normal nuclear translocation<br />

after melphalan stimulation compared to unstimulated and TNF-stimulated controls (data<br />

not shown). Analogous results have been found when NFκB translocation was<br />

measured using the Cellomics HitKit NFκB Activation Assay (Pittsburgh, PA, USA)<br />

(figure 27C).<br />

To summarize these data, NFκB seemed to translocate to the nucleus following<br />

melphalan treatment, but inhibition of NFκB p65 translocation into the nucleus had<br />

neither enhanced toxic effects only at very low concentrations of melphalan, nor any<br />

beneficial influence on the cellular outcome.<br />

4.13.2 Role of p38 MAPK<br />

p38 MAPK is another signaling molecule which has been shown to be involved in<br />

various signaling cascades. P38 MAPK among others is also activated following TNFR1<br />

stimulation with TNF. It has been shown to activate ATF which is able to translocate to<br />

the nucleus. Pre-treatment of primary liver cells with 1 µM of a specific inhibitor for p38<br />

MAPK and incubated with melphalan did also not influence the toxicity induced by<br />

melphalan as determined by cell viability assay (figure 28 A). Measurement of caspase<br />

activity (figure 28 B) confirmed this result, indicating that melphalan-induced toxicity was<br />

also independent of p38 MAPK signaling.<br />

- 96 -


4.13.3 Role of JNK<br />

Results<br />

A third important cellular signaling pathway is related to a kinase called Junk<br />

kinase (JNK). JNK has been shown to be activated though MADD and MEK 4/7 and<br />

leads to nuclear translocation of c-Jun upon activation. C-Jun in turn is a transcriptional<br />

activator that led to transcription of TNF, IL-1ß and CD95L mRNA. Initial experiments<br />

using JNK Inhibitor II from Calbiochem (Darmstadt, Germany) revealed that this inhibitor<br />

was toxic itself in higher concentrations (7-10 µM) and that the optimum range was<br />

between 4 µM and 6 µM which was in line with other data obtained in our laboratory<br />

(Georg Dünstl, personal communication). Incubations performed on primary murine<br />

hepatocytes with melphalan and JNK-Inhibitor II pre-treatment (30 min, 5 µM), displayed<br />

that cell death induced by melphalan was significantly reduced (p < 0.001), but not<br />

totally blocked (figure 29 A). Median values calculated for that specific sets of<br />

experiments revealed an EC50 for melphalan of about 90 µg/ml, whereas an EC50 of<br />

about 175 µg/ml was calculated for melphalan in cells pre-treated with JNK-Inhibitor II<br />

(30 min, 5 µM) (figure 29). The beneficial effect of JNK inhibition in melphalan-induced<br />

cytotoxicity was also observed by light microscopy, showing an intact cellular shape of<br />

hepatocytes up to 24 hours, with some apoptotic cells also observed in between roughly<br />

estimated to 15 - 25%. To investigate if this inhibitory effect could also observed in<br />

- 97 -


Results<br />

incubations performed with melphalan and TNF in the absence of Kupffer cells, we<br />

prepared purified hepatocytes and incubated these cells with melphalan and TNF with<br />

and without JNK-Inhibitor II pre-treatment (30 min, 5 µM). Interestingly, the protective<br />

effect of JNK inhibition was also be observed. EC50 values were calculated at about 75<br />

µg/ml for melphalan/TNF alone versus 220 µg/ml under presence of JNK-Inhibitor II<br />

(figure 29), indicating that the beneficial effect of JNK inhibition was due to effects on the<br />

hepatocyte and not related to Kupffer cell actions.<br />

4.13.4 Role of PARP<br />

The idea of the so-called “Papirmeister-Hypothesis” is that a huge activation of<br />

PARP resulting from massive nicks and breaks in the DNA might lead to a drop in<br />

cellular NAD + -pool and therefore might contribute to apoptosis induction. Since<br />

melphalan was shown to induce single- and double strand breaks in the DNA, we were<br />

also interested, if this hypothesis might contribute to melphalan-induced hepatocyte<br />

death.<br />

Therefore, we investigated, if PARP-activation could be measured in isolated<br />

primary hepatocytes or Kupffer cells upon stimulation with genotoxic agents. To address<br />

this problem, we prepared purified Kupffer cells and hepatocytes, stimulated them with<br />

H2O2 or MNNG in presence or absence of the specific PARP inhibitor PJ-34 and<br />

- 98 -


Results<br />

measured PARP activity using the Cellomics Array Scan (Pittsburgh, PA, USA). PARPactivation<br />

could be detected in isolated Kupffer cells (data not shown) as well as in<br />

hepatocytes (figure 30A) following H2O2 or MNNG treatment. Activation of PARP could<br />

be blocked by co-incubation with PARP inhibitor PJ-34, indicating a specific<br />

measurement of PARP. Interestingly, we found only a weak activation of PARP in<br />

melphalan-treated cells, which could be also blocked by PJ-34 (data not shown). To<br />

address this problem in a different setting, we incubated primary murine liver cells in<br />

presence and absence of various concentrations of PJ-34 (ranging from 1.5 µM up to<br />

100 µM) with melphalan. But also in this setting no influence on melphalan-induced<br />

toxicity was measured, stressing previous results (figure 30B, data only shown for 100<br />

µM). EC50 values were calculated in that specific set of experiments at 60 µg/ml for<br />

melphalan versus 70 µg/ml for melphalan co-incubated with PJ-34 (figure 30B).<br />

In summary, a weak activation of PARP was detected following melphalan<br />

incubation, but this slight activity played no decisive role in melphalan-induced<br />

hepatocyte death.<br />

- 99 -


Results<br />

4.14 Influence of melphalan on cellular redox and energy state and vice<br />

versa<br />

In previous projects in our laboratories the influence of cellular redox and energy<br />

state on hepatocyte apoptosis was investigated. Briefly, it could be demonstrated that<br />

depletion of glutathione inhibited receptor-mediated liver apoptosis in vivo (298). Under<br />

ATP-depleted conditions it could be demonstrated that TNFR-mediated apoptosis is<br />

blocked in vitro and in vivo in contrary to CD95-mediated apoptosis (305). Since<br />

previous results indicated that melphalan-induced toxicity is strongly dependent on TNF,<br />

TNFR1 and TNFR2 we were interested if modulation of cellular redox or energy state<br />

might influence cellular outcome in response to stimulation with melphalan.<br />

4.14.1 Melphalan and cellular redox state<br />

Melphalan is known to react spontaneously with glutathione also without the<br />

activity of glutathione-S-transferase. Since it was previously shown that melphalanrelated<br />

toxicity was enhanced in patients following intracellular GSH-depletion and the<br />

cellular redox state influences the cellular decision to die or not to die and how to die<br />

(apoptotic or necrotic) we investigated if cellular redox state is relevant for melphalaninduced<br />

cell death or if the pool of cellular glutathione is influenced by the drug.<br />

To address these questions, we isolated primary hepatocytes of tnfr1r2 genedeficient<br />

mice, purified them as described above to avoid melphalan induced cellular<br />

toxicity and incubated them with different concentrations of melphalan. As shown in<br />

figure 31A for concentrations ranging from 50 to 200 µg/ml melphalan the drug slightly<br />

modifies cellular glutathione, but as compared to controls no significant depletion of GSX<br />

could be observed after 18 hours. Interestingly, when cellular glutathione was depleted<br />

by adding diethylmaleate (DEM) in a final concentration of 1 mM to primary liver cell<br />

culture 30 min before melphalan incubation, melphalan-induced cellular toxicity was<br />

significantly enhanced at concentrations higher than 25 µg/ml as compared to<br />

incubations without DEM (figure 31B). To verify these data in the isolated perfused<br />

mouse liver, we prepared livers of mice depleted of glutathione by an intraperitoneal<br />

injection of 500 mg/kg phorone dissolved in 300 µl of corn oil and perfused them with<br />

150 mg/kg melphalan. Intrahepatic depletion of glutathione also enhanced melphalan<br />

toxicity, whereas depletion of glutathione alone by phorone application had no influence<br />

on perfusion parameters and liver injury as indicated by lacking ALT activity in the<br />

perfusate (figure 31C).<br />

- 100 -


Results<br />

In conclusion these results are in line with data obtained in the cell culture system<br />

and negate ideas of apoptosis protection by glutathione depletion in this model.<br />

4.14.2 Melphalan and cellular energy status<br />

Initial experiments demonstrated melphalan induced toxicity being strongly<br />

dependent on TNF and TNF receptors leading to hepatocyte apoptosis. Since our group<br />

- 101 -


Results<br />

was previously able to demonstrate that TNF-mediated hepatic apoptosis is blocked<br />

when primary hepatocytes or mice were pre-treated with ATP-depleting carbohydrates<br />

and cellular ATP content subsequently was depleted to about 20%. Therefore, we<br />

investigated, if melphalan-induced hepatocyte apoptosis is also influenced by reduced<br />

cellular energy state due to incubation with ATP-depleting carbohydrates.<br />

To address this idea we isolated primary liver cells from C57Bl/6 mice, preincubated<br />

them with 25 mM of ATP-depleting carbohydrates such as fructose, tagatose<br />

or sorbitol for 30 min and added melphalan in usual concentrations to the cells.<br />

Interestingly, under ATP-depleted conditions melphalan-induced cell death was<br />

significantly decreased (figure 32A-C). To investigate if this effect was due to cellular<br />

ATP depletion or possible other side-effects, e.g. osmolarity, we incubated liver cells<br />

derived from wild type mice with different concentrations of melphalan in presence of the<br />

non-ATP-depleting carbohydrates such as glucose, mannose or sorbitose (25 mM, t=-30<br />

min) (figure 32D-F).<br />

In that specific subsets of experiments we could show that cells pre-incubated<br />

- 102 -


Results<br />

With ATP-depleting carbohydrates were significantly protected from melphalan-induced<br />

apoptosis, while cells pre-treated with non-ATP-depleting sugars were still sensitive for<br />

melphalan. With these experiments we could demonstrate that ATP depletion protects<br />

primary livers cells also when both TNF receptors are involved in apoptosis induction.<br />

Since we could show that hepatocyte apoptosis can also be induced by melphalan<br />

and exogenous TNF in absence of Kupffer cells, we were further interested, if<br />

melphalan/solTNF-induced apoptosis can also be blocked by ATP depletion.<br />

Therefore, we purified murine hepatocytes incubated them with 25 mM of fructose<br />

for 30 min and stimulated them with melphalan and soluble TNF. As shown in figure 33,<br />

cells were also protected significantly for apoptosis induced by melphalan and TNF.<br />

- 103 -


Results<br />

At this point it should be noted, that with respect to the high numbers of animal<br />

experiments these experiments have not been verified in cells lacking TNFR1 or TNFR2<br />

or both since it has been shown sufficiently that under Kupffer cell depleted conditions<br />

hepatocyte death following melphalan/TNF treatment is due to activation of TNFR1 and<br />

hepatocytes derived of animals lacking TNFR1 are not dying in response to<br />

melphalan/TNF treatment.<br />

Of greater importance for this project was the role of ATP depletion in the isolated<br />

perfused mouse liver. Previous experiments in animals showed that the intrahepatic<br />

ATP content can be lowered to about 20% 30 min after an intraperitoneal injection of 5<br />

g/kg of tagatose, to about 40% 30 min after an intraperitoneal injection of 5 g/kg of<br />

fructose and to about 30% 30 min after an intraperitoneal injection of 10 g/kg of fructose.<br />

Each of these intracellular ATP concentrations has been shown to be sufficient to block<br />

hepatocyte apoptosis in vivo due to TNF-mediated liver injury (Latta et al., personal<br />

communication). To ensure sufficient ATP-depletion, mice were injected 30 min before<br />

liver preparation with doses of 10 g/kg fructose or glucose as a control sugar. 30 min<br />

after performing a closed recirculating fashion, freshly dissolved melphalan was added<br />

in a dose of 150 mg/kg to the perfusate and perfusate samples were taken as indicated.<br />

As expected, livers of mice injected with fructose were completely protected from<br />

melphalan-induced organ-disruption up to nine hours (figure 34 ○), whereas livers of<br />

mice pre-treated with glucose were still sensitive showing a mean ALT-release of 3,000<br />

(± 1,874) U/L in the perfusate after 360 min (figure 34 ◊). Untreated livers from mice<br />

depleted of ATP displayed only a slight toxicity after nine hours of perfusion, but this low<br />

release of ALT might be also due to this long time of perfusion, since also untreated<br />

control livers displayed slightly increased ALT-concentrations in the perfusate after nine<br />

to ten hours (data not shown).<br />

- 104 -


Results<br />

To sum these data up, as expected from pervious data melphalan-induced<br />

hepatocyte death could be inhibited by depletion of the intrahepatic ATP-pool. This<br />

finding is of further interest, because protection of TNFR-mediated apoptosis could only<br />

be demonstrated for primary cells, whole organs or living animals and not for cells or cell<br />

lines derived from carcinomas. Therefore, it would be interesting to investigate if livers of<br />

tumor-bearing mice show selective apoptosis of tumor cells when perfused with<br />

melphalan under ATP-depleted conditions. We addressed this question and the results<br />

obtained in that experiments are shown in the next section.<br />

- 105 -


Results<br />

4.15 Effect of melphalan on DEN-induced liver tumors in isolated liver<br />

perfusion of tumor-bearing mice<br />

ATP depletion by carbohydrates was only possible in primary liver cells and not in<br />

cell lines derived from hepatic carcinoma. In addition, Adelman et al. published that<br />

fructokinase activity was lacking in rat liver tumors (310). If this is the case, this<br />

difference between normal healthy hepatic tissue and transformed hepatic carcinoma<br />

cells might be useful for cancer therapy.<br />

4.15.1 DEN induced massive tumor growth in murine livers following<br />

intraperitoneal injection<br />

To address the question if normal healthy liver tissue in contrary to tumorigenic<br />

tissue might be protected from melphalan-induced apoptosis under ATP-depleting<br />

conditions we performed tumor studies in C57Bl/6 mice induced by i.p. injection of DEN<br />

as described in materials & methods. Tumors were grown for at least 20 weeks and<br />

finally mice were sacrificed and tumor-bearing livers were perfused with melphalan<br />

and/or TNF in the presence and absence of ATP-depleting carbohydrate fructose.<br />

Mice injected with DEN displayed enhanced neoplasm in the liver (figure 35). The<br />

tumors varied in size and numbers from small singular pre-neoplastic lesions (figure<br />

35A+B) to solid well vascularised tumors spread over the whole organ (figure 35D-F).<br />

Interestingly, perfusion parameters did not differ between the individual livers being<br />

more or less affected by tumor growth. Additionally, untreated livers from tumor-bearing<br />

mice perfused only with buffer did not show altered perfusion parameters compared to<br />

untreated control livers of normal C57Bl/6 mice.<br />

- 106 -


Results<br />

- 107 -


Results<br />

4.15.2 Isolated liver perfusion of tumor-bearing livers with melphalan/TNF<br />

When perfused with melphalan alone livers displayed enhanced ALT release in the<br />

perfusate, decreasing perfusate flow though the organ and finally breakdown of capillary<br />

flow about 360 min following melphalan application (figure 36A ▲). On the contrary,<br />

when perfused with melphalan under ATP-depleted conditions only a minor ALT release<br />

could be detected till perfusion was stopped after 480 min not differing significantly from<br />

control livers (figure 36A ♦, ○). Additionally, perfusate flow was stable till the end of<br />

experiment. When perfused with melphalan and TNF, livers displayed enhanced ALT<br />

release in the perfusate as compared to untreated control livers, resulting in breakdown<br />

of perfusate flow about 390 min after melphalan application (figure 36B ▲). As<br />

compared to livers perfused with melphalan alone perfusate flow was slightly enhanced<br />

and the breakdown of capillary perfusate flow was also delayed. In Addition, there was<br />

less ALT release detectable in the perfusate (figure 36A+B ▲). When perfused<br />

simultaneously with melphalan and TNF under ATP-depleted conditions livers displayed<br />

enhanced ALT release in the perfusate compared to untreated control livers, but<br />

decreased and delayed ALT-release compared to livers perfused with melphalan and<br />

TNF in the presence of cellular ATP (figure 36B).<br />

- 108 -


Results<br />

In conclusion, ATP depletion was not sufficient to protect organs completely from<br />

melphalan/TNF-induced toxicity in this setting as this could be demonstrated in IPML of<br />

livers derived from healthy animals. If this enhanced ALT release of ATP-depleted livers<br />

was related to tumor cell death could not be estimated based on ALT activity in the<br />

perfusate. This problem was further investigated by measurement of caspase-activation<br />

in liver homogenates and histological examination of liver slices collected after the<br />

experiment.<br />

4.15.3 Activation of caspases in normal and neoplastic tissue<br />

In order to investigate if healthy and tumor tissue was affected by melphalan or<br />

melphalan combined with TNF, we analysed liver samples of normal and neoplastic<br />

tissue collected after the experiment. Liver samples isolated from healthy tissue<br />

displayed no significant onset of caspase activity (ctrl: 51±27; melphalan: 110±87;<br />

melphalan/fructose: 88±42; melphalan/TNF 217±284; melphalan/TNF/fructose 90±93<br />

pg/mlxmin). Additionally, also caspase activity detected in homogenates of isolated<br />

neoplastic tissue differed not significantly between the individual treatment groups<br />

(melphalan: 51±29; melphalan/fructose: 171±193; melphalan/TNF 335±413;<br />

melphalan/TNF/fructose 100±109 pg/mlxmin). According to this results, it was not<br />

possible conclude anything concerning the treatment-related onset of apoptosis in<br />

normal and neoplatic tissue coout of the results obtained.<br />

Therefore, we tried to answer the question if ATP depletion might contribute to a<br />

selective death of cancer cells by histological examination of liver slices prepared from<br />

each perfused mouse liver.<br />

4.15.4 Histological examination of perfused livers derived from tumor-bearing<br />

mice<br />

Liver slices from perfused mouse livers were stained with H.E. and TUNEL to have<br />

an selective marker for apoptosis. In order to grade tissue damage in these liver slices<br />

we scored specimens on a scale from zero to five with zero being not damaged at all<br />

and five being heavily damaged with total disruption of liver architecture and<br />

apoptosis/necrosis been observed in most areas. Untreated control livers displayed an<br />

intact liver architecture indicating that no or at least very minor toxicity occurred due to<br />

- 109 -


Results<br />

the perfusion conditions (score 0; figure 37E+F). Livers treated with melphalan alone<br />

displayed apoptosis and necrosis in tumors as well as in normal tissue. Interestingly,<br />

observed morphological changes varied between the individuals in this treatment group,<br />

ranging from very minor to heavy tissue destruction (score 1-4, mostly score 2, figure<br />

37A). But in all specimens observed toxicity was comparable between tumor and normal<br />

tissue.<br />

When melphalan was perfused in combination with TNF a similar situation could be<br />

observed, but the level of toxicity was much higher (score 5, figure 37C). Toxicity<br />

observed in liver samples of this treatment group matched very well with the early flow<br />

stop seen during perfusion and the massive ALT release observed. But differences<br />

between normal and tumor tissue could hardly be seen.<br />

One of the most interesting questions was how ATP depletion would influence cell<br />

death in normal and tumor tissue. According to the perfusion parameters, livers should<br />

be protected in case of melphalan and partially protected in case of melphalan/TNF<br />

treatment. Interestingly, this was not the case. Livers perfused with melphalan alone<br />

under ATP-depleted conditions displayed also some light to moderate toxicity (score 2-3,<br />

figure 37B), but compared to the non-ATP-depleted condition induced toxicity by<br />

melphalan seemed to be decreased. Differences between neoplastic and normal liver<br />

tissue could also not be observed.<br />

In organs perfused with melphalan and TNF under ATP-depleted conditions, toxicity was<br />

also observed in normal and tumor tissue, but tumor tissue seemed to be more affected<br />

(tumor tissue score 3-4; normal tissue score 1-2, figure 37D). As compared to perfusions<br />

under non-ATP-depleted conditions, the onset of cell death in tumor tissue was not<br />

altered, but the normal surrounding tissue seemed to be more intact.<br />

Concerning the mode of cell death it has to be mentioned that apoptosis as well<br />

as necrosis could be observed. Especially in the presence of melphalan and TNF<br />

necrosis was be found in tumor tissue, also affecting the vascular system. If this related<br />

to TNF is not clear.<br />

In conclusion, a clear statement if APT-depletion protects normal tissue from<br />

melphalan-induced liver damage whilst keeping tumor tissue sensitive for induction of<br />

cell death can not be made. Interestingly, when having a closer look on individual<br />

experiments, results obtained in histological assessment of liver slices can be matched<br />

to singular results of caspase activity and are also in line with perfusion parameters. But<br />

- 110 -


Results<br />

compared to other individual experiments of the same treatment schedule, differences in<br />

the severity of tissue destruction became obvious.<br />

- 111 -


5 Discussion<br />

Discussion<br />

Non-resectable cancers confined to the liver still represent a significant clinical<br />

problem. In most of these cases, the liver is not the primary site of malignancies, but<br />

represents the most susceptible site for metastasis. More than 140,000 new patients are<br />

diagnosed with colorectal cancer each year in the United States and approximately more<br />

than 25% of them develop liver metastases, of which most of them are non-resectable<br />

(42, 408). Median survival of these patients is poor, ranging from 12 to 18 months, even<br />

with local or systemic chemotherapy (158, 409-411). In most of these cases,<br />

chemotherapy or a combination of surgery and chemotherapy is the only way of<br />

therapeutic intervention. Since systemic application of therapeutics has been failed for<br />

cancer therapy due to severe toxicity mainly affecting the gastrointestinal tract and the<br />

immune system, several techniques have been developed for regional treatment of<br />

cancers confined to various organs (46, 62, 91, 412, 413). For the liver, mainly two<br />

techniques have been developed, which are commonly used in therapy today: isolated<br />

hepatic perfusion (IHP) and hepatic artery infusion (HAI) (51, 174). These elegant<br />

procedures have been shown to prevent systemic toxicity, due to leakage of applied<br />

alkylating agents, or cytokines such as TNF in the circulation, and as such did not allow<br />

systemic toxicity. However, the hepatic perfusion of melphalan, especially in<br />

combination with TNF, has been shown to lead to a reversible hepatotoxicity in the<br />

majority of patients, by means of a mechanism that remains elusive.<br />

In this study, the effect of melphalan in primary murine hepatocytes, as well as in<br />

isolated mouse liver perfusion was investigated to identify possible mechanisms leading<br />

to cell death induced by this alkylating agent. The purpose of this study was to elucidate<br />

mechanisms responsible for adverse reactions to melphalan, based on observations<br />

gathered from numerous clinical studies and discussed in several reviews for being<br />

responsible for enhanced adverse reactions. Last but not least, the knowledge<br />

accumulated on death-receptors in the last decades opened up some new possibilities<br />

of how to intervene with receptor-mediated apoptosis and might thus be interesting for<br />

reducing side-effects related to cancer therapy.<br />

- 131 -


Discussion<br />

5.1 The role of TNF and TNF receptors in melphalan-mediated<br />

hepatotoxicity<br />

Hepatotoxicity, as well as hypotension, have been shown before to represent<br />

major side-effects associated with the systemic use of TNF in cancer therapy (414).<br />

Although TNF and CD95 ligand represent the key cytokines implicated in hepatotoxicity<br />

and hepatocyte apoptosis. The former cytokine, the expression of which is upregulated<br />

in many acute and chronic liver diseases, does not cause apoptosis in hepatocytes in<br />

vitro or in vivo, unless during conditions of ischemia (415) or transcriptional/translational<br />

inhibition (416). In contrast, the interaction between TNF and its TNFR1 has been shown<br />

to be crucial in the priming of hepatocytes during liver regeneration (198, 417, 418) and<br />

in the proliferation of oval cells during the neoplastic phase of liver carcinogenesis (419).<br />

Interestingly, recent results have indicated a crucial role of TNF in classical toxicological<br />

processes (420). Indeed, TNF was shown to be implicated in the regulation of products<br />

inducing inflammation and fibrosis, but not in direct hepatocyte damage in carbon<br />

tetrachloride hepatotoxicity (421). Also in alcohol-mediated liver toxicity, an important<br />

role for TNF was recently suggested (422). Moreover, the interaction between TNF and<br />

TNFR2 was suggested to be implicated in fumonisin hepatotoxicity in mice (423).<br />

In this study, evidence for a novel mechanism is presented by which an<br />

antineoplastic drug induces a significant hepatotoxicity in mice, mediated by membranebound<br />

TNF on the surface of Kupffer cells. The importance of TNF in the observed<br />

melphalan cytotoxicity is stressed by means of several observations: 1) Hepatocytes<br />

lacking either TNFR1 or TNFR2 both TNF receptors were protected from melphalan<br />

cytotoxicity. 2) Co-incubation of melphalan and TNF-neutralising anti-TNF antibodies<br />

totally abolished melphalan-mediated hepatocellular apoptosis. 3) In absence of Kupffer<br />

cells, identified as the source of TNF production, no hepatocyte apoptosis could be<br />

detected. Last but not least, 4) lacking toxicity of macrophage-depleted hepatocytes<br />

could either be restored by adding isolated Kupffer cells or by adding soluble TNF to<br />

cellular incubations containing melphalan. With that set of experiments all four<br />

necessary conditions have been investigated to identify TNF as the toxic mediator of<br />

hepatocyte death following melphalan incubation: 1) loss of the TNF receptors abolished<br />

melphalan toxicity (receptor-knock-out experiments); 2) neutralisation of the ligand<br />

prevented melphalan toxicity (experiments with neutralising antibodies); 3) the source of<br />

TNF was identified (experiments in absence of Kupffer cells); 4) Toxicity could be<br />

restored when either the mediator or the natural source of the mediator was added to<br />

the system only including hepatocytes and melphalan (TNF or Kupffer cells added to<br />

purified hepatocytes in the presence of melphalan).<br />

- 113 -


Discussion<br />

Although TNFR1, that contains a death domain (218, 223), has been shown to<br />

mediate apoptosis induced by soluble TNF in hepatocytes (212, 416), not only this<br />

receptor type, but also TNFR2 was causally implicated in the melphalan-induced<br />

hepatotoxicity, since livers from both tnfr1 and tnfr2 gene-deficient mice were resistant.<br />

This means that although TNFR2 does not contain a death domain, it is nevertheless<br />

implicated in melphalan-induced apoptosis. It was previously shown that this TNF<br />

receptor type, as well as membrane-bound TNF were also implicated in experimental<br />

liver injury caused by the plant lectin Con A, in inflammation and degeneration<br />

processes in the central nervous system of transgenic mice (424-426) as well as in<br />

experimental cerebral malaria (373). Moreover, transmembrane TNF was found to be<br />

sufficient to mediate localised tissue toxicity and chronic inflammatory arthritis in<br />

transgenic mice (426), as well as in concanavalin A hepatotoxicity (248). Since<br />

membrane-bound TNF, induced by melphalan, has been suggested to preferentially<br />

trigger TNFR2 and soluble TNF preferentially activates TNFR1, this could explain why<br />

melphalan and soluble TNF had an additive effect in hepatotoxicity, since the activation<br />

of TNFR2 has been reported to enhance the TNFR1-mediated apoptosis (reviewed in<br />

(427)). Indeed, the stimulation of TNFR2 was recently shown to lead to the degradation<br />

of TRAF2 subsequently blocking TNFR1-mediated apoptosis (260). Co-stimulation or<br />

pre-stimulation of TNFR2 was shown to enhance caspase-8 processing, while TNFR2<br />

was suggested to compete with TNFR1 for the recruitment of newly synthesised TRAF2bound<br />

anti-apoptotic factors, thereby promoting the formation of a caspase-8-activating<br />

TNFR1 complex (261). A similar co-operative signaling of both TNF receptors triggered<br />

by transmembrane TNF was also reported in a murine model of arthritis (428).<br />

Furthermore, studies with mutated TNF demonstrated that TNF mutants with a<br />

selective binding capacity for TNFR1 displayed similar potency to wild type TNF in<br />

causing cytotoxicity of a human laryngeal carcinoma-derived cell line (HEp-2) and<br />

cytostasis in a human leukaemia cell line (U937), while exhibiting lower pro-inflammatory<br />

activity than wild type TNF (252). Nevertheless, the role of TNFR2 in the cytocidal effect<br />

of TNF still is a matter of debate. A study in TNF sensitive cells using broad panel of<br />

antibodies against extracellular domains of TNFR2 suggests that TNFR2 contributes to<br />

the cytocidal effect of TNF both by its own signaling and by regulating the access of TNF<br />

to TNFR1 (242). This is also strengthened by studies demonstrating a severe<br />

inflammatory syndrome involving mainly the pancreas, liver, kidney, and lung in<br />

transgenic mice, resulting from production of the human p75TNF-R in relevant levels, as<br />

compared to human diseases. In addition, this process was shown to evolve<br />

independently of the presence of TNF, lymphotoxin alpha, or the TNFR1 (429).<br />

Additionally it was found that overexpression of TNFR2 lead to TNF-dependent<br />

apoptosis in transfected HeLa cells (257). Last but not least Grell et al. could<br />

demonstrate in a TNF cytotoxicity model of KYM-1-derived cell line that both receptors<br />

- 114 -


Discussion<br />

are able to induce apoptosis, as revealed from a similar onset of DNA fragmentation and<br />

typical morphologic criteria. This provides strong evidence for a distinct signaling<br />

pathway of TNFR1 and TNFR2, indicating protein kinase(s)-mediated control of TNFR1<br />

signaling and a tight linkage of TNFR2 to arachidonate metabolism (430).<br />

But cell death is not the only cellular process involving TNFR1 and TNFR2 cooporation<br />

(431-433). It could also be shown that ligand-independent TNFR55-mediated<br />

co-operation takes place in TNFR75-induced granulocyte/macrophage-CSF secretion. In<br />

contrast to ligand-dependent apoptosis in PC60 cells this co-operation is unidirectional.<br />

According to the results of Declercq et al. this ligand-independent co-operation is<br />

crucially dependent on an intact TNFR-associated factors 1 and 2 (TRAF1/TRAF2)binding<br />

domain, while dominant negative TRAF2 has no inhibitory effect (258).<br />

5.2 Poseidon’s trident - mechanism of melphalan-induced cytotoxicity<br />

TNF was clearly identified as being the toxic principle of melphalan-induced<br />

hepatocyte apoptosis by triggering death receptor TNFR1. But according to previous<br />

findings on TNF-mediated apoptosis a number of questions remained open. As<br />

documented in several publications, TNF alone failed in apoptosis induction in vitro or in<br />

vivo unless co-administration of translational or transcriptional inhibitors such as alphaamanitin,<br />

actinomycin D or the aminosugar galactosamine (416). Since in our<br />

experimental setting none of these inhibitors have been added, and soluble TNF has<br />

been induced apoptosis in purified hepatocytes in the presence but not in the absence of<br />

melphalan, this indicated that melphalan simultaneously blocked transcription in liver<br />

cells as has been demonstrated in HeLa and melanoma cells before (434, 435). On the<br />

other hand, melphalan was shown to cross-link DNA in vitro (436) and is therefore likely<br />

to be sufficient to block DNA transcription of anti-apoptotic proteins. So, melphalan fulfils<br />

at least two jobs: stimulation of Kupffer cells and blocking transcription of cellular DNA.<br />

In addition to these findings, another interesting observation was made. While<br />

being unable to detect released soluble TNF upon incubations with melphalan, we<br />

proposed a role for membrane-bound TNF that has also been shown to be biologically<br />

active. Membrane-bound TNF mediates its biological action mainly through TNFR2<br />

since this receptor has a much higher affinity for membrane-bound than for soluble TNF<br />

(437). The striking role for membrane-bound TNF could be demonstrated when both<br />

cell-types, hepatocytes and Kupffer cells, were co-incubated but spatially separated by<br />

cell culture inserts allowing soluble TNF to pass through. In this case, no apoptosis was<br />

induced in hepatocytes.<br />

- 115 -


Discussion<br />

The second result stressing this hypothesis was the observation that the study<br />

drug was able to completely inhibit LPS-induced TNF production by Kupffer cells.<br />

Moreover, when administered to LPS-challenged mice, also no TNF release could be<br />

detected 90 min following LPS injection. This cytokine release was also inhibited when<br />

melphalan was administered up to one hour following LPS challenge, demonstrating that<br />

lack of soluble TNF was not due to transcriptional inhibition but due to inhibition of<br />

cytokine release. Since soluble TNF has been shown to be cleaved from membranebound<br />

TNF by means of the activity of the TNF converting enzyme (71, 72, 438) this<br />

lead to the hypothesis that melphalan inhibits TACE activity. Indeed, we clearly<br />

demonstrated that melphalan inhibited recombinant TACE activity in a cell-free system.<br />

Interestingly, IC50 calculated from the obtained results fitted very well to results obtained<br />

in cell culture experiments. The onset of hepatocyte apoptosis was observed at a<br />

concentration of 25 to 50 µg/ml of melphalan, being exactly the concentration at which<br />

the activity of the TACE enzyme was blocked in vitro.<br />

However, there exist also observations arguing against this hypothesis: First,<br />

melphalan inhibits both LPS-induced secreted and membrane-bound TNF production in<br />

Kupffer cells, whereas TACE-inhibitors should only inhibit the former and second,<br />

melphalan-induced membrane-bound TNF disappears already after 60 min, which would<br />

not be the case upon inhibition of TACE. However, these findings might also be<br />

explained by the fact that melphalan blocks efficiently transcription and translation<br />

preventing cellular reactions due to external or internal stimuli already on DNA or mRNA<br />

level while enhanced membrane bound TNF expression on the cellular surface in the<br />

first two hours might result from presentation of already existing intracellular pools of<br />

pre-formed TNF as suggested by Gordon et al. for mouse peritoneal mast cells releasing<br />

their content upon IgE-dependent activation (79).<br />

Melphalan on the one hand induced hepatotoxicity mediated via TNF and TNF<br />

receptors in cell culture, but on the other hand never produced measurable amounts of<br />

soluble TNF upon stimulation of isolated Kupffer cells or in the perfusate of isolated<br />

livers treated with different doses of melphalan. Moreover, melphalan seemed to<br />

abrogate TNF secretion of isolated Kupffer cells, peritoneal macrophages or isolated<br />

livers upon stimulation with various concentrations of endotoxin. An inhibitory effect of<br />

melphalan on the ELISA protocol has been excluded, since TNF has been detected by<br />

the ELISA in samples containing melphalan and spiked with TNF. In our working<br />

hypothesis we proposed that melphalan abolished TNF release by blocking the<br />

matrixmetalloproteinase (MMP) TACE, cutting TNF off the membrane. The inhibition of<br />

TACE by melphalan could finally be proven in a cell-free assay since the recombinant<br />

protein and various fluorogenic substrates and inhibitors have been made commercially<br />

available. Further more, we were also able to demonstrate lacking TNF release in vivo<br />

- 116 -


Discussion<br />

since mice challenged with endotoxin, SEB, or the plant lectin Con A were protected<br />

from induced liver failure when pre-treated with 50 to 100 mg/kg of melphalan, Indicating<br />

that this ability of TACE inhibition might also takes place in the living animal. In this<br />

series of animal experiments we were also able to block endotoxin-induced TNF release<br />

when melphalan was given up to one hour following LPS injection, arguing against the<br />

fact that lack of TNF is only due to transcriptional or translational inhibition (439). Similar<br />

findings were published by Nakama et al. for GalN/LPS induced liver injury. In this study,<br />

the topoisomerase II inhibitor etoposide was shown to protect mice from GalN/LPS<br />

induced liver failure when administered 50, 26 and 4 hours before GalN/LPS challenge<br />

(440). Interestingly, this etoposide effect was also demonstrated in GalN/TNF-mediated<br />

liver injury, which could be explained by upregulation of Bcl-xL, an anti-apoptotic protein<br />

in the liver, in response to etoposide-treatment. In our model, we could clearly<br />

demonstrate that low or undetectable cytokine levels in sera of mice occurred when<br />

melphalan was co-administered, which is in contrast to the study mentioned above.<br />

But TNF was not the only cytokine being reduced in response to melphalan.<br />

Additionally, also IFNγ was reduced in GalN/SEB, Con A, and αCD95 challenged mice<br />

and IL-1ß was found to be reduced in mice challenged with Con A. For the Con A model<br />

it was described before that injection of this lectin caused increased plasma<br />

concentrations of TNF, IL-6, IL-1, IL-2, and IFNγ, while TNF was identified as the central<br />

mediator of Con A-induced liver injury (393, 441). Another study on Con A-induced<br />

hepatotoxicity that investigated the role of TNF suggested Kupffer cells as the major<br />

source of TNF produced in the liver (382). In general, the cross talk between<br />

macrophages and T cells mostly mediated by IFNγ was found to be important for Con A<br />

toxicity to occur (382), while TNF was identified as the cytokine being responsible for<br />

triggering apoptosis in the hepatocyte. Therefore, our data are in line with these previous<br />

findings. Interestingly, in the in vitro system the 26 kD membrane-bound form of TNF<br />

was found to be crucial for the Con A induced toxicity, less the soluble 17 kD form (442),<br />

while in vivo soluble TNF contributed to toxicity (394, 395, 441, 443). If this dependence<br />

of both TNF receptors and the involvement of membrane-bound TNF in hepatocyte<br />

apoptosis in vitro also contributes to the distinct morphological alterations observed in<br />

both models remains elusive.<br />

- 117 -


Discussion<br />

5.3 Melphalan-induced upregulation of membrane TNF<br />

In view of these results another question had to be answered. Melphalan was<br />

shown to block TNF release probably by inhibition of matrix-metalloproteinase TACE in<br />

vitro and in vivo. On the other hand, melphalan was shown to induce transcriptional<br />

arrest in cells. The question that remained to be answered was whether melphalan also<br />

induced an upregulation of TNF by gene activation or if only already existing pools of<br />

pre-formed TNF were presented similar to findings mentioned by Gordon et al. (79).<br />

Evidence for melphalan up-regulating TNF comes from observations published for<br />

MOPC-315 tumor-bearing mice. Jovasevic and Mokyr could demonstrate that one hour<br />

after low-dose melphalan application to mice increased mRNA levels for IFN-ß could be<br />

detected in the spleen (352). IFN-ß in turn were shown to be necessary for upregulation<br />

of TNF expression in MOPC-315 tumor cells. According to results obtained in cell-based<br />

ELISA, a slight upregulation of TNF was also observed in the first two hours following<br />

melphalan incubation, which was also confirmed in western blot analysis of RAW<br />

stimulated with melphalan.<br />

But also a series of previous results and recent findings argued against an active<br />

upregulation of TNF. First of all, melphalan has been shown to react very rapidly with<br />

DNA (436), disabling cellular response mechanisms involving transcription and<br />

translation. This might explain why the melphalan-induced expression of membrane TNF<br />

in Kupffer cells takes place only at early time points and subsequently fades out. This<br />

implies that presented TNF originates from pre-formed mRNA pools present in Kupffer<br />

cells, as reported by others (444). Second, pre-treatment of primary liver cells with<br />

various phosphodiesterase-inhibitors one hour before adding melphalan did not show<br />

any alteration of melphalan-induced TNF-mediated hepatocyte apoptosis (data not<br />

shown). PDE-inhibitors have been shown to effectively block TNF upregulation of<br />

isolated Kupffer cells in response to LPS-treatment. Third, atrial natriuretic peptide<br />

(ANP), which has also been shown to effectively block TNF production of Kupffer cells in<br />

response to various pathologic situations such as ischemia reperfusion injury in rat (445)<br />

and LPS-triggered primary macrophages and RAW cells in vitro (445, 446), had also no<br />

influence on melphalan-induced hepatotoxicity, stressing the fact that no active<br />

upregulation of TNF takes place. On the other hand, western blot analysis of memTNF<br />

revealed that two hours following melphalan incubation an enhanced band of<br />

approximately 26 kD could be detected, confirming results previously obtained by cellbased<br />

ELISA. Nevertheless, this might also be due to processing of already pre-formed<br />

RNA pools as suggested by Ulich et al. (444).<br />

In summary, it seems unlikely that an active upregulation of TNF in response to<br />

melphalan takes place in our model system but this might also be a matter of melphalan<br />

- 118 -


Discussion<br />

concentrations used and pharmacological parameters like uptake and metabolism and<br />

local cellular concentration.<br />

5.4 Clinical relevance<br />

Our findings might contribute to the fact why tumor therapy or local tumor therapy<br />

is more efficient when melphalan is co-administered together with TNF, as compared to<br />

treatment strategies using melphalan or TNF alone as revealed by several clinical<br />

studies (65, 174, 447-449). On the other hand, results of one pre-clinical study<br />

describing a lack of anti-tumor activity of human recombinant TNF, alone or in<br />

combination with melphalan in a nude mouse human melanoma xenograft system<br />

cannot be explained (450). On the contrary, in our model system apoptosis of primary<br />

murine hepatocytes was also induced by melphalan and recombinant human TNF with<br />

concentrations and kinetics comparable to murine TNF. These results are also<br />

supported by findings of Leist et al. demonstrating TNFR1-mediated apoptosis of<br />

primary murine hepatocytes also in response to incubation with recombinant human<br />

TNF (212).<br />

Apart from cellular results, the high extent of hepatotoxicity we observed in our<br />

isolated perfused mouse liver model does not correspond to the relatively mild<br />

hepatotoxicity reported in advanced studies in patients treated with melphalan and TNF<br />

via hepatic perfusion. One of the reasons for this could be that this treatment procedure<br />

frequently uses hyperthermia, i.e. a condition which could give rise to increased levels of<br />

heat shock proteins (HSP). Recent results suggested that induction of heat shock<br />

protein 70 in mice kept at 42°C, can protect them from the systemic toxicity of TNF,<br />

without interfering with the tumoristatic effect in a murine melanoma tumor model (451).<br />

Since we show in our pre-clinical setting that melphalan hepatotoxicity is mediated by<br />

TNF, hyperthermia could thus potentially confer protection under clinical conditions.<br />

Moreover, our observation that the melphalan-inducible membrane-bound TNF fades<br />

out already 2 hours after melphalan treatment, could correlate with the reversible nature<br />

of the melphalan-linked hepatotoxicity observed in patients.<br />

To summarize this part, we could clearly demonstrate that TNF is the toxic<br />

principle of melphalan-induced hepatocyte toxicity involving both TNF receptors.<br />

Additionally, we could proof that there is a striking role for Kupffer cells presenting<br />

membrane-bound TNF on their surface and that physical interaction of both cell types<br />

(hepatocytes and Kupffer cells) is necessary to transmit TNF action to the hepatocyte,<br />

while an active upregulation of TNF, as compared to LPS-treated macrophages of livers<br />

seems unlikely or at least needs further investigation. Last but not least, we could<br />

- 119 -


Discussion<br />

demonstrate that melphalan is able to block TACE activity in vitro resulting in lack of<br />

soluble TNF upon Kupffer cell stimulation in all models investigated and abundant liver<br />

toxicity in various models of experimental liver failure in vivo.<br />

5.5 Differential role of TNF receptors in the presence and absence of Kupffer<br />

cells<br />

According to numerous articles and reviews published on the role of TNF<br />

receptors in apoptosis, it seems to be proven, that TNFR1 alone is sufficient to initiate a<br />

cascade of biochemical reactions in response to binding of the specific ligand (192, 212,<br />

230, 430). This interaction finally leads to activation of caspases and subsequently<br />

cellular breakdown, ending up in apoptosis. Therefore it is interesting that also<br />

numerous articles have been published demonstrating a role for both TNF receptors in<br />

various experimental settings or pathological situations as mentioned above.<br />

Interestingly, in most of these models a crucial role for TNFR2 was only demonstrated in<br />

cells or animals being deficient of this receptor. The exact mechanism how TNFR2 is<br />

involved in the mediated toxicity is currently not available.<br />

Since membrane-bound TNF, induced by melphalan, has been suggested to<br />

preferentially trigger TNFR2 and soluble TNF preferentially activates TNFR1, this could<br />

explain why melphalan-mediated toxicity is dependent of TNFR2. According to this, the<br />

activation of TNFR2 was reported to enhance the TNFR1 mediated apoptosis (reviewed<br />

in (427)). Another study undertaken in 3T3 cells further indicated an alternative role for<br />

TNFR2. This study revealed that although TNFR2 can significantly reduce the TNF<br />

concentration required for cell killing, the mechanism by which this is accomplished is<br />

not through the generation of an intracellular signal by TNFR2. Instead, TNFR2<br />

regulates the rate of TNF association with TNFR1, possibly by increasing the local<br />

concentration of TNF at the cell surface through rapid ligand association and<br />

dissociation (255). This phenomenon is also known as ligand passing in literature.<br />

As compared to findings published by Li et al, concerning TRAF2 degradation in<br />

response to TNFR2 triggering (260) and Alexopoulou et al. concerning the positive cooperation<br />

between both TNF receptors in murine arthritis (428) mentioned above, a<br />

similar co-operation might occur in our model system.<br />

When cells or organs depleted of Kupffer cells were pre-treated with melphalan<br />

and stimulated with exogenous soluble TNF, the onset of apoptosis occured more<br />

rapidly but with comparable extent. In this setting TNFR1 is the only necessary receptor,<br />

clearly demonstrating that TNFR2 is not necessary for intracellular caspase activation. In<br />

- 120 -


Discussion<br />

contrast to the studies of Alexopoulou et al. mentioned above we did not observe a<br />

delayed onset of injury when TNFR2 is missing, in our model system the toxicity was<br />

totally abolished. This might be the case when TACE is inhibited by melphalan, so that<br />

the cells were unable to release soluble TNF and transmembrane TNF was not<br />

presented in sufficient amounts to trigger TNFR1 as demonstrated by western blot and<br />

cell-based ELISA analysis.<br />

This means that at least two distinct mechanisms exist. In the presence of Kupffer<br />

cells and absence of exogenous TNF melphalan toxicity is mediated by Kupffer cellpresented<br />

transmembrane TNF and both TNF receptors, while in the absence of Kupffer<br />

cells and the presence of exogenous TNF only the presence of TNFR1 is a necessary<br />

condition for inducing hepatocyte death (for illustration see figure 38). Interestingly, our<br />

study also supports findings of Gorelik et al. who could demonstrate a striking role for<br />

TNF production for the curative effectiveness of low-dose melphalan for mice bearing a<br />

large MOPC-315 tumor (448).<br />

Taken together these results might explain why melphalan in combination with<br />

exogenous soluble TNF is more efficient in tumor therapy leading to higher regression<br />

and response rates, as compared to the treatment strategies using melphalan alone. On<br />

the other hand the presented data also fit to findings of higher liver toxicity observed in<br />

patients post-surgery and combinatorial treatment strategies published by several<br />

authors (172, 447, 452, 453) and therefore contribute to a better understanding of the<br />

underlying mechanism.<br />

- 121 -


Discussion<br />

5.6 Intracellular signaling of TNF receptors following melphalan<br />

incubation<br />

Binding of TNF to its specific receptors is followed by a whole variety of different<br />

cellular responses in various cell types. TNF was shown to be a key mediator in<br />

inflammation and sepsis, immunological reactions, apoptosis, rheumatoid arthritis, liver<br />

regeneration and so on (69, 70, 86-90). To fulfil these tasks by interaction with only two<br />

receptors, there have to be other intracellular control centres regulating cellular<br />

decisions upon TNF receptor activation. In the last decades numerous molecules,<br />

pathways and transcription factors have been identified being involved in the regulation<br />

of cellular responses upon TNF stimulation. In case of liver regeneration reactive oxygen<br />

species, glutathione content, NFκB, STAT3 and AP1 have all been identified as major<br />

regulators (reviewed in (198)), It was shown that autoregulation of tnf-alpha gene<br />

transcription by selective signaling through TNFR1 requires p38 mitogen-activated<br />

protein (MAP) kinase activity and the binding of the transcription factors ATF-2 and Jun<br />

to the TNF-alpha cAMP-response element (CRE) promoter element. Consistent with<br />

- 122 -


Discussion<br />

these findings, TNFR1 engagement results in increased p38 MAP kinase activity and<br />

p38-dependent phosphorylation of ATF-2 (454). Studies in mice deficient in NFκB<br />

subunits have shown that this transcription factor is important for lymphocyte responses<br />

to antigens and cytokine-inducible gene expression. In particular, the RelA (p65) subunit<br />

is required for induction of tumor necrosis factor-alpha (TNF-alpha)-dependent genes<br />

(455). HeLa cells transfected with TNFR2 could be shown to display enhanced<br />

activation of NFκB and c-Jun kinase (257). Therefore, we investigated at least three of<br />

these molecules that have been demonstrated to be involved in TNF-mediated<br />

pathways: NFκB, MAPK and c-Jun kinase.<br />

5.6.1 Nuclear factor κ B<br />

Most normal and neoplastic cell types are resistant to TNF cytotoxicity unless<br />

when co-treated with protein- or RNA-synthesis inhibitors, such as cycloheximide and<br />

actinomycin D. Cellular resistance to TNF requires TNF receptor-associated factor 2<br />

(TRAF2), which has been hypothesised to act mainly by mediating activation of the<br />

transcription factors NFkB and activator protein 1 (AP1) (456). Especially NFκB p65 has<br />

been demonstrated to be at least partly responsible for this effect by inducing a set of<br />

genes upon TNF receptor activation that downregulate the apoptosis signal (457). In<br />

case of TNFR1 triggering the intracellular adapter protein TRADD, which in turn interacts<br />

with TRAF2, has been identified, leading to two distinct cellular responses, activation of<br />

cell death and activation of NFκB (221, 222). Another group could demonstrate that<br />

recruitment of the signal transducer FADD to the TNFR1 complex mediates apoptosis<br />

but not NFκB activation. Two other signal transducers, RIP and TRAF2, instead mediate<br />

NFκB activation (279). These two responses, however, seem to diverge downstream to<br />

TRAF2 which might be mediated by MAP3K-related kinase NIK (266).<br />

For TNFR2, also an ability of NFκB activation was reported. Of special interest<br />

are results reporting the identification of a novel TNFR2 receptor isoform termed<br />

icp75TNFR, which is generated by the use of an alternative transcriptional start site<br />

within the TNFR2 gene and characterised by regulated intracellular expression. The<br />

icp75TNFR binds to TNF and mediates intracellular signaling. Overexpression of the<br />

icp75TNFR cDNA results in TNF-induced activation of NFκB in a TRAF2-dependent<br />

manner (458).<br />

Interestingly, these observations published on TNF receptor-mediated NFκB<br />

activation are in line with findings in our model system. Since neither inhibitors of NFκB<br />

nor cells/organs derived from inducible NFκB-k.o. mice have been displayed<br />

morphological or biochemical changes in apoptosis induction in response to melphalan,<br />

nor NFκB p65 translocation to the nucleus has not been altered by melphalan, we have<br />

- 123 -


Discussion<br />

to conclude that NFκB signaling does not play a decisive if any role in melphalan<br />

mediated toxicity. This holds also true for incubations with melphalan and soluble TNF in<br />

the absence of Kupffer cells. Since we have been demonstrate that TNFR2 has not<br />

been involved in that specific setting a striking role of NFκB in the cross talk between<br />

TNFR1 and TNFR2 under Kupffer cell-containing conditions can be excluded.<br />

5.6.2 p38 mitogen activated protein kinase and c-Jun N-terminal kinase<br />

The response of cells to extracellular stimuli is mediated in part by a number of<br />

intracellular kinase and phosphatase enzymes. Within this area of research the<br />

activation of mitogen-activated protein (MAP) kinases have been extensively described<br />

and characterised as central components of the signal transduction pathways stimulated<br />

by both growth factors and G protein-coupled receptor agonists. Signaling events<br />

mediated by these kinases are fundamental to cellular functions, such as proliferation,<br />

differentiation and cell death. More recently, homologues of the p42 and p44 isoforms of<br />

MAP kinase have been described namely the stress-activated protein kinases (SAPKs)<br />

or alternatively the c-jun N-terminal kinases (JNKs) and p38 MAP kinase. These MAP<br />

kinase homologues are integral components of parallel MAP kinase cascades activated<br />

in response to a number of cellular stresses including inflammatory cytokines (e.g.<br />

Interleukin-1 (IL-1) and TNF), heat and chemical shock, bacterial endotoxin and<br />

ischemia/cellular ATP depletion. These MAP kinase homologues mediate the<br />

transduction of extracellular signals to the nucleus and are pivotal in the regulation of the<br />

transcription events that determine functional outcome in response to such stresses<br />

(reviewed in (459)). For renal ischemia reperfusion injury it has been shown that TNF<br />

was released from the kidney in response to oxidants originating from renal ischemia<br />

reperfusion injury. These oxidants in turn activated p38 MAP kinase and the TNF<br />

transcription factor NFκB leading subsequently to TNF synthesis. In a positive feedback,<br />

pro-inflammatory fashion, binding of TNF to specific TNF membrane receptors can<br />

reactivate NFkB. This provides a mechanism by which TNF can upregulate its own<br />

expression as well as facilitate the expression of other genes pivotal to the inflammatory<br />

response (reviewed in (460, 461)). MAPKs have also been shown to take part in the<br />

signal transduction of LPS-stimulated cell activation to produce bioactive substances<br />

including TNF, while p38 MAPK is an important regulator of TNF expression. MAPKs<br />

also take part in the complex network of signals induced by ionising radiation and other<br />

cellular stresses, such as TNF and endoplasmatic reticulum (ER) stress (reviewed in<br />

(462, 463)).<br />

In our model system, inhibition of p38 MAPK had no significant influence on<br />

melphalan-induced TNF-mediated hepatocyte apoptosis with caspase activity also not<br />

- 124 -


Discussion<br />

being altered. This fits very well into the concept that melphalan toxicity is not involving<br />

an active upregulation of TNF, as suggested by data obtained from western blot and<br />

cell-based ELISA experiments. Also positive feedback mechanisms as discussed for<br />

pro-inflammatory events can be excluded. This is also in line with observations that<br />

melphalan-induced Kupffer cell-mediated toxicity is not due to LPS contaminations of the<br />

study drug, as excluded by Limulus amoebocyte lysate (LAL) bioassay measurements.<br />

In view of these data it was surprising that inhibition of JNK displayed a protective<br />

effect on melphalan-induced hepatocyte death in the presence as well as in the absence<br />

of Kupffer cells. JNK is a stress-activated protein kinase that can be induced by<br />

inflammatory cytokines, bacterial endotoxin, osmotic shock, UV radiation, and hypoxia.<br />

Inhibition of JNK phosphorylation has been shown to prevent expression of the<br />

inflammatory genes encoding for COX-2, IL-2, IFN-gamma, TNF-alpha, and prevented<br />

the activation and differentiation of primary human CD4 cell cultures. In animal studies,<br />

JNK inhibition blocked (bacterial) lipopolysaccharide-induced expression of tumor<br />

necrosis factor-alpha and inhibited anti-CD3-induced apoptosis of CD4(+) CD8(+)<br />

thymocytes (464). Studies on bile acid biosynthesis in cultured rat hepatocytes indicated<br />

that the JNK pathway plays a pivotal role in regulating CYP7A1 levels, the rate-limiting<br />

enzyme in the neutral pathway of bile acid biosynthesis (465). Further evidence for an<br />

implication of JNK was shown in, fibroblast-like synoviocytes and synovium during<br />

inflammation and tissue destruction in rheumatoid arthritis (466), in ischemia reperfusion<br />

injury in mouse liver (467 , 468) and in apoptosis induced by DNA damage due to<br />

stimulation with etoposide, teniposide, and UV irradiation (345). In the latter model, it<br />

turned out that the stress-activated kinase pathway (SAPK/JNK) was required for the<br />

maximal induction of apoptosis. If this would also be the case in melphalan-induced<br />

hepatocyte death, this might explain why inhibition of JNK at least leads to a reduction in<br />

hepatocyte death following melphalan incubation. This finding is also in line with the<br />

observation that inhibition of JNK affects both mechanisms of melphalan-induced cell<br />

death, i.e. those occurring either in the presence or in the absence of Kupffer cells.<br />

Interestingly, TRAF2 was been demonstrated to be required for TNF-mediated activation<br />

of c-Jun N-terminal kinase, but binding of TNF to TNFR2 induced ubiquitinylation and<br />

proteasomal degradation of TRAF2 (260). As mentioned above, cellular resistance to<br />

TNF required TRAF2. Inhibition of TRAF2 function(s) by signaling-deficient<br />

oligomerisation partners or by molecules affecting TRAF2 recruitment to the TNFR1<br />

complex completely abrogated the cytoprotective response. (456).<br />

In conclusion this means that if TRAF2 has been degraded in response to TNFR2<br />

stimulation, the cellular resistance to TNF has been lost and apoptosis has been<br />

promoted. This could partially explain why TNFR2 is crucially involved in melphalanmediated<br />

cell death since membrane TNF has been shown to play a crucial role in this<br />

scenario and the latter has been demonstrated to bind preferentially to TNFR2. But this<br />

- 125 -


Discussion<br />

scenario would not explain why partial protection of hepatocytes from cell death occurs<br />

by means of JNK-inhibition. Instead this partial protection could be explained, assuming<br />

that TRAF2 is not degraded completely and remaining amounts being sufficient to<br />

activate JNK. However, this was not investigated and remains speculative.<br />

5.7 Modification of melphalan-toxicity by altered cellular redox or energy<br />

state<br />

5.7.1 Glutathione<br />

The major low molecular weight thiol inside cells, the tripeptide glutathione (GSH),<br />

is of importance for protection of the cell against oxidative challenge, for thiol<br />

homeostasis required to guarantee basic functions, and for defence mechanisms<br />

against xenobiotics (reviewed in (296)). In the last decades numerous examples of toxic<br />

liver injury have been published involving intracellular depletion of GSX caused by<br />

carbon tetrachloride, acetaminophen overdoses or allylalcohol demonstrating the<br />

importance of an intact glutathione status (294). Intracellular depletion of GSH as such<br />

has been shown to have little metabolic consequences unless an additional stress is<br />

superimposed. The kinetic properties of GSH-dependent enzymes imply that loss of up<br />

to 90% of intracellular GSH may still be compatible with cellular integrity. Mitochondrial<br />

GSH accounting for about 10% of total cellular GSH may define the threshold beyond<br />

which toxicity commences. Such a severe depletion of GSH has been described for<br />

some diseases such as liver dysfunction, AIDS (295) or pulmonary fibrosis. In case of<br />

receptor-mediated apoptosis, protection from cell death could be recently demonstrated<br />

in case of CD95-mediated apoptosis and TNF receptor-mediated apoptosis (297).<br />

Interestingly, this protective effect could only be shown for apoptosis while necrosis was<br />

shown to be enhanced (298). Recently the underlying mechanism of this cytoprotective<br />

effect could be identified, since the processing of pro-caspase-8 activation at the DISC<br />

has been demonstrated to be dependent of an intact GSX-status (469). In another<br />

cellular model, glutathione depletion has been shown to be associated with decreased<br />

Bcl-2 expression and increased apoptosis (303).<br />

Also in cancer therapy the effect of glutathione during drug treatment has been<br />

intensively studied. For melphalan it could be shown that this molecule interacts very<br />

rapidly with glutathione, also without the need of GSX-S-transferase (470). A study<br />

performed on the rate of glutathione conjugation of melphalan demonstrated significant<br />

- 126 -


Discussion<br />

amounts of conjugation also in the absence of cellular enzymes which could not be<br />

enhanced by addition of melanoma cell homogenates (181).<br />

Based on the knowledge accumulated on the role of glutathione on receptormediated<br />

apoptosis we also investigated the role of glutathione depletion in melphalaninduced<br />

hepatocellular apoptosis. Cells or organs depleted of glutathione by DEM or<br />

phorone displayed enhanced cytotoxicity, when incubated with even low concentrations<br />

of melphalan, as compared to non-depleted cells or organs. These findings are in line<br />

with enhanced gastrointestinal toxicity in mice observed upon treatment with regional<br />

hyperthermia and melphalan following BSO-mediated glutathione depletion (176-178)<br />

and studies obtained in U-937 human promonocytic cells (471). Conjugation of cytostatic<br />

drugs to glutathione was also discussed as a potential mechanism of tumor cell<br />

resistance (472), but this was negated since data from rats and humans have been<br />

suggested that hepatic GSH conjugation plays a very minor (if any) role in the<br />

elimination of melphalan. Therefore, it seems unlikely that modulation of GSH levels<br />

affects the rate of elimination of this drug (182). This is also in line with our results<br />

demonstrating that addition of melphalan in concentrations of up to 200 µg/ml to isolated<br />

purified hepatocytes is not sufficient to deplete intracellular GSX.<br />

In conclusion, our data obtained with melphalan under glutathione-depleted<br />

conditions fit in very well with observations in cell-free and cellular studies as well as<br />

with pre-clinical and clinical studies undertaken with melphalan. All experimental<br />

systems demonstrated that melphalan is able to conjugate spontaneously to glutathione<br />

and depletion of glutathione is leading to enhanced cytotoxicity. Therefore, depletion of<br />

GSX is most likely not suitable for cancer therapy.<br />

5.7.2 ATP<br />

The cellular ATP content is another apoptosis-related parameter (473, 474) that<br />

may be modified selectively in the liver. ATP depletion has been found to affect at least<br />

three different pathways. First, there is evidence that ATP is needed in the formation of<br />

the so-called apoptosome in which pro-caspase-9 is activated (238, 239, 475).<br />

Accordingly, cellular ATP depletion frequently prevented activation of execution<br />

caspases downstream of cytochrome c release (476, 477). Such cells with damaged<br />

mitochondria are always destined to die, but when ATP is lacking, a switch from<br />

apoptosis to necrosis has been observed (476-478). Second, in some models the ATPdependent<br />

step might be upstream of cytochrome c release (476, 477), and third in<br />

CD95-mediated apoptosis being independent of the mitochondrial pathway, ATP<br />

depletion has no effect on caspase activation (473) or cell death (479). In studies on<br />

- 127 -


Discussion<br />

receptor-mediated apoptosis we could previously demonstrate, that TNF-induced liver<br />

cell injury was completely blocked upon ATP-depletion, while CD95-mediated<br />

hepatocyte apoptosis was enhanced (305). Interestingly, this could also been<br />

demonstrated in vivo since ATP depletion induced by pre-treatment of mice by singular<br />

i.p. injection of ATP-depleting carbohydrates has been prevented TNF-mediated liver<br />

injury (305).<br />

Since we have been demonstrated, that melphalan-induced hepatocyte death is<br />

also dependent of TNF and TNF receptors, we investigated the effect of ATP-depleting<br />

and non-ATP-depleting sugars in our model system. Analogous to previous data on Act<br />

D/TNF-induced hepatocyte apoptosis we could demonstrate a protective effect of the<br />

ATP-depleting sugars fructose and tagatose in the melphalan-model, while non-ATPdepleting<br />

sugars like glucose, mannose or sorbose displayed no protective effect.<br />

Interestingly, this cytoprotective effect could only be demonstrated for primary liver cells<br />

and not for HepG2 cells and Hepa1,6 cells (data not shown), which is in line with<br />

findings previously published for transformed cells or cells derived from hepatoma (305),<br />

as well as primary neurones (480), and lymphoid cells (474). These results further<br />

indicate that melphalan-induced apoptosis is only dependent of TNF and TNF receptor<br />

mediated apoptosis, since CD95-induced apoptosis was shown to be enhanced upon<br />

lack of intracellular ATP. Also of interest is the fact that ATP depletion protects cells or<br />

organs also in a model which is dependent of both TNF receptors. This could previously<br />

be demonstrated only for Con A-induced murine liver failure in vivo (Latta et al., personal<br />

communication).<br />

In this thesis evidence is also provided that isolated murine livers can be<br />

protected from TNF-mediated or melphalan-induced liver injury by ATP depletion. This is<br />

of greater interest, since this model has been closely related to the situation of patients<br />

undergoing IHP during cancer therapy. The cytoprotective effect of ATP-depleting<br />

carbohydrates could be demonstrated by a lacking efflux of intra-hepatic ALT in the<br />

perfusate, by lacking caspase-activity in liver samples and in histological examination<br />

demonstrating an intact liver morphology. Moreover, perfusion parameters like perfusate<br />

flow through the liver were observed to be improved, at least after several hours of<br />

perfusion.<br />

- 128 -


Discussion<br />

5.8 Influence of ATP depletion on melphalan/TNF-mediated organ toxicity<br />

in tumor-bearing mice<br />

In this study, we provide evidence that normal hepatocytes in contrast to<br />

transformed neoplastic cells might be protected by ATP depletion from melphalaninduced<br />

apoptosis. This idea was investigated in studies performed in isolated perfusion<br />

of livers derived from tumor-bearing mice. Mice injected intraperitoneally with DEN<br />

displayed massive neoplastic lesions spread over the whole liver. The treatment of mice<br />

with DEN has been shown to produce solid, well-vascularised tumors about 20 weeks<br />

following i.p. injection. Tumor induction in our study varied from small pre-neoplastic<br />

lesions to huge tumors spread over the whole organ. As expected, in the absence of<br />

TNF melphalan induced liver toxicity in tumor-bearing mice was comparable to livers<br />

isolated from healthy mice concerning the amount and the time-course of ALT release in<br />

the perfusate. Additionally, pre-treatment with fructose also protected hepatocytes from<br />

melphalan-induced toxicity. These data might reflect observations in patients undergoing<br />

IHP only being treated with melphalan alone. In these patients also low tumor response<br />

rates and low hepatocyte toxicity have been observed (54, 57). When livers were<br />

perfused in the presence of TNF, the observed liver toxicity was much higher and ATP<br />

depletion by fructose pre-treatment was not able to prevent organ toxicity completely.<br />

But this was also expected, because tumor cells should undergo apoptosis or necrosis<br />

in this setting. When compared to the situation in the patient, high liver toxicity or<br />

fulminant organ failure leading to death was only observed when cytostatic drugs were<br />

combined with exogenous soluble TNF for therapy (65, 172, 174, 447). Since it could not<br />

be determined whether ALT release in the perfusate resulted from tumor cells or healthy<br />

hepatocytes we tried to measure caspase activity in isolated healthy liver tissue and<br />

tumor tissue after the experiment. Unfortunately, this approach failed because resulting<br />

values of caspase activities were not consistent in three independent assays.<br />

Histological examination of liver slices collected after the experiment also revealed no<br />

unified results. Interestingly, untreated control livers displayed less to no toxicity with an<br />

overall intact liver architecture, indicating that perfusion conditions alone did not<br />

influence tumor cell or hepatocyte death. This is also in line with previous findings of<br />

liver slices collected from healthy mice perfused as controls for up to nine hours.<br />

Interestingly, when tumor-bearing livers were perfused with melphalan, no obvious<br />

differences could be observed between tumor tissue and healthy cells in most cases. In<br />

some material, necrosis of the endothelium could be observed, but if this was due to<br />

melphalan treatment or the mode of perfusion is not clear since this necrosis has been<br />

also observed in some livers perfused with melphalan under ATP-depleted conditions.<br />

When livers were depleted of ATP before melphalan perfusion, we also could not<br />

observe an unified situation in all individuals of the same treatment group. Some of the<br />

material displayed no apoptosis at all, some livers displayed apoptosis in healthy tissue<br />

- 129 -


Discussion<br />

as well as at the tumor site and in some material we could find enhanced cytotoxicity in<br />

tumors, as compared to healthy tissue. Interestingly, this was either observed in mice<br />

bearing huge tumors spread over the whole lobe or in one case of very low tumor<br />

induction, bearing only very small lesions with diameters smaller than 1 mm. The latter<br />

might be explained by the observation that very small lesions in the liver are only<br />

connected to the hepatic artery (45) and therefore might not be very well perfused when<br />

livers were connected via the hepatic veins. When livers were perfused in combination<br />

with melphalan and TNF, much higher toxicity could be observed. This is also in line with<br />

observed reduced perfusate flow though the liver following 4 hours of perfusion. But<br />

whether this contributed to the extended toxicity, for example through changes in the<br />

micro-circulation, is not clear. Again, mice pre-treated with fructose for ATP depletion in<br />

this treatment group displayed slightly enhanced toxicity at the tumor site, as compared<br />

to the healthy tissue.<br />

In summary, according to the results obtained from the isolated liver perfusion<br />

with melphalan/TNF under ATP-depleted and non-depleted conditions, pre-treatment of<br />

livers/mice with fructose seemed to have a positive effect on the reduction of unwanted<br />

side-effects by melphalan/TNF on unaffected liver tissue. But when having a closer look<br />

by histological examination this selective protective effect between normal healthy tissue<br />

and transformed tumor tissue is less clear. To obtain more significant results larger<br />

studies have to be performed, also with various concentrations of used drugs and<br />

cytokines.<br />

In general, the understanding of the mechanism and time-course of melphalan<br />

hepatotoxicity in the presence and the absence of exogenous TNF can thus provide a<br />

basis for the design of clinical treatment regimens that minimise or even abrogate the<br />

hitherto inevitable adverse effects of this therapy.<br />

- 130 -


6 Summary<br />

Summary<br />

Isolated hepatic perfusion of non-resectable liver cancer using the combination of<br />

hyperthermia, tumor necrosis factor (TNF) and melphalan can be associated with a<br />

treatment-related hepatotoxicity. The present thesis investigated whether, apart from<br />

TNF, also melphalan is cytotoxic in primary murine liver cells in vitro. Additionally, it was<br />

investigated whether further mediators are involved, which cell types contribute to the<br />

cytotoxicity and which mode of cell death takes place. The elucidation of the mechanism<br />

of melphalan-induced liver cell destruction as proposed here may allow a rational<br />

therapy of patients suffering from inoperable hepatic tumors or metastases.<br />

In detail the following results were obtained:<br />

1. Melphalan induced a concentration-dependent apoptosis in isolated murine liver<br />

cells in vitro and isolated perfused mouse livers in situ.<br />

2. Cells or organs derived from TNF receptor 1, TNF receptor 2 and TNF receptor 1<br />

and 2 gene-deficient mice, isolated primary cells or organs of mice depleted of<br />

Kupffer cells were resistant against melphalan-induced toxicity. Incubation of cells<br />

with melphalan in presence of neutralising antiTNF antibodies abrogated<br />

melphalan-induced apoptosis.<br />

3. Direct contact of hepatocytes and Kupffer cells was mandatory for the observed<br />

toxicity.<br />

4. In absence of Kupffer cells, melphalan-induced toxicity was restored in the<br />

presence of exogenous recombinant TNF dependent only on TNF receptor 1.<br />

5. The activity of recombinant TNF alpha converting enzyme (TACE) was shown to<br />

be inhibited by melphalan. In situ and in vivo, the production of soluble TNF upon<br />

LPS-stimulation and hepatotoxicity upon treatment of mice by various proinflammatory<br />

stimuli was blocked by melphalan.<br />

6. After pre-treatment of cells or mice with ATP-depleting carbohydrates such as<br />

fructose or tagatose, primary hepatocytes and isolated livers were protected from<br />

melphalan-induced apoptosis in contrast to non-ATP-depleting sugars such as<br />

glucose, mannose or sorbose. On the contrary, depletion of intracellular<br />

- 131 -


Summary<br />

glutathione by diethylmalate in vitro or phorone in the isolated liver perfusion in<br />

situ enhanced melphalan-induced apoptosis.<br />

7. Transformed cells or cell lines derived from hepatoma cells such as Hepa1-6 or<br />

HepG2 cells were not protected by ATP-depletion. This fact offers the possibility<br />

to selectively and transiently protect healthy cells against melphalan in the ATPdepleted<br />

state.<br />

8. Such a strategy was tested using an ex-vivo approach of tumor-bearing mice in<br />

isolated liver perfusion in the ATP-depleted state and preliminary data were<br />

obtained which support the concept.<br />

- 132 -


7 Zusammenfassung<br />

Zusammenfassung<br />

Isolierte hepatische Perfusion mit Melphalan, Tumor Nekrose Faktor (TNF) und<br />

Hyperthermie zur Behandlung von Patienten mit nicht-operablen Lebertumoren ist oft<br />

mit akuter Hepatotoxizität assoziiert. In der vorliegenden Arbeit wurde die Frage<br />

untersucht, ob Melphalan unabhängig von TNF toxisch auf primäre murine Leberzellen<br />

wirkt. Zusätzlich wurde die Art der Schädigung, sowie beteilige Mediatoren und<br />

Zelltypen untersucht. Die Aufklärung des Mechanismus, auf dem die Leberschädigung<br />

induziert durch Melphalan beruht, ermöglicht eine bessere Therapie von Patienten mit<br />

inoperablen Tumoren oder Metastasen in der Leber.<br />

Im Einzelnen wurden folgende Ergebnisse erhalten:<br />

1. Melphalan induzierte konzentrationsabhängig Apoptose in isolierten primären<br />

Leberzellen in vitro und in isolierten Organperfusion in situ<br />

2. Zellen oder Organe, die aus Mäusen mit einem genetischen Defekt im TNF<br />

Rezeptor 1, TNF Rezeptor 2 oder beiden TNF Rezeptoren isoliert wurden,<br />

waren vor Melphalan induzierter Schädigung geschützt. Organe, in denen die<br />

Kupfferzellen experimentell depletiert wurden, zeigten ebenfalls keine<br />

Melphalan-induzierte Schädigung. Coinkubation von Melphalan mit neutralisierenden<br />

anti-TNF Antikörpern verhinderte die Melphalan-induzierte<br />

Schädigung ebenso.<br />

3. Ein direkter Kontakt zwischen Hepatozyt und Kupfferzelle war notwendig, um<br />

die Melphalan-induzierte Schädigung zu vermitteln.<br />

4. In Abwesenheit von Kupfferzellen konnte die Melphalan-induzierte<br />

Schädigung durch Zugabe von exogenem rekombinantem TNF wieder<br />

hergestellt werden. In diesem Falle war die induzierte Schädigung nur von<br />

TNF Rezeptor 1 abhängig.<br />

5. Die Aktivität von rekombinanter TNF-Convertase (TACE) wurde durch<br />

Melphalan inhibiert. In situ und in vivo wurde die Inhibition von TACE durch<br />

die ausbleibende TNF-Freisetzung nach Stimulation mit Endotoxin und durch<br />

den Schutz vor TNF-vermittelter Leberschädigung nach Behandlung von<br />

Mäusen mit pro-inflammatorischen Substanzen gezeigt.<br />

- 133-


Zusammenfassung<br />

6. Vorbehandlung von Zellen oder Mäusen mit ATP-depletierenden Zuckern wie<br />

Fruktose oder Tagatose schützte primäre Hepatozyten oder isoliert<br />

perfundierte Lebern vor Melphalan-induzierter Schädigung, während die<br />

Vorbehandlung mit nicht-depletierenden Zuckern wie Glukose, Mannose oder<br />

Sorbose keinen Einfluß hatte. Im Gegensatz dazu hatte eine Vorbehandlung<br />

zur Depletion des intrazellulären Glutathionspiegels mit Diethylmaleat in vitro<br />

oder Phoron in situ keinen protektiven Effekt, sondern erhöhte die Schädigung<br />

durch Melphalan.<br />

7. Transformierte Zellen oder Zelllinien, die aus Lebertumoren kultiviert wurden<br />

wie Hepa 1-6 oder HepG2 Zellen, waren nicht durch ATP depletion geschützt.<br />

Diese Tatsache bietet die Möglichkeit, gesunde Hepatozyten unter ATP<br />

depletion selektiv und transient gegen Melphalan zu schützen.<br />

8. Eine derartige Strategie wurde in einem ex-vivo Ansatz untersucht, in dem<br />

Tumor-induzierte Mäuse in der isolierten Leberperfusion mit Melphalan und<br />

TNF unter ATP-depletierten Bedingungen behandelt wurden. Vorläufige Daten<br />

aus dieser Studie bestätigten grundsätzlich das Konzept.<br />

- 134 -


8 References<br />

References<br />

1. Mattioli A. Kolonial- oder Vernichtungskrieg? Entgrenzte Kriegsgewalt - Der<br />

italienische Giftgaseinsatz in Abessinien 1935-1936. Vierteljahreshefte für<br />

Zeitgeschichte 2003;3:311-338.<br />

2. Adair CPJ, Bagg H. Experimental and clinical studies on the treatment of cancer<br />

by dichlorethylsulfide (mustard gas). Ann Surg 1931;93:190-199.<br />

3. Gilman A, Philips FS. The biological actions and therapeutic applications of the<br />

B-chloroethyl amines and sulfides. Science 1946;103:409-415.<br />

4. Chen JJ, Sun Y, Nabel GJ. Regulation of the proinflammatory effects of Fas<br />

ligand (CD95L). Science 1998;282:1714-1717.<br />

5. Price CC, Gaucher GM, Koneru P, Shibakawa R, Sowa JR, Yamaguchi M.<br />

Relative reactivities for monofunctional nitrogen mustard alkylation of nucleic acid<br />

components. Biochim Biophys Acta 1968;166:327-359.<br />

6. Lawley PD, Brookes P. Acidic dissociation of 7:9-dialkylguanines and its possible<br />

relation to mutagenic properties of alkylating agents. Nature 1961;192:1081-<br />

1082.<br />

7. Bauer GB, Povirk LF. Specificity and kinetics of interstrand and intrastrand<br />

bifunctional alkylation by nitrogen mustards at a G-G-C sequence. Nucleic Acids<br />

Res 1997;25:1211-1218.<br />

8. Henne T, Schmahl D. Occurrence of second primary malignancies in man--a<br />

second look. Cancer Treat Rev 1985;12:77-94.<br />

9. Wang P, Bauer GB, Bennett RA, Povirk LF. Thermolabile adenine adducts and<br />

A.T base pair substitutions induced by nitrogen mustard analogues in an SV40based<br />

shuttle plasmid. Biochemistry 1991;30:11515-11521.<br />

10. Wang P, Bauer GB, Kellogg GE, Abraham DJ, Povirk LF. Effect of distamycin on<br />

chlorambucil-induced mutagenesis in pZ189: evidence of a role for minor groove<br />

alkylation at adenine N-3. Mutagenesis 1994;9:133-139.<br />

11. Meier HL, Gross CL, Papirmeister B. 2,2'-Dichlorodiethyl sulfide (sulfur mustard)<br />

decreases NAD+ levels in human leukocytes. Toxicol Lett 1987;39:109-122.<br />

12. Szabo C, Dawson VL. Role of poly(ADP-ribose) synthetase in inflammation and<br />

ischaemia- reperfusion. Trends Pharmacol Sci 1998;19:287-298.<br />

13. de Murcia G, Menissier de Murcia J. Poly(ADP-ribose) polymerase: a molecular<br />

nick-sensor. Trends Biochem Sci 1994;19:172-176.<br />

14. Lautier D, Lagueux J, Thibodeau J, Menard L, Poirier GG. Molecular and<br />

biochemical features of poly (ADP-ribose) metabolism. Mol Cell Biochem<br />

1993;122:171-193.<br />

- 135-


References<br />

15. Sims JL, Berger SJ, Berger NA. Poly(ADP-ribose) Polymerase inhibitors preserve<br />

nicotinamide adenine dinucleotide and adenosine 5'-triphosphate pools in DNAdamaged<br />

cells: mechanism of stimulation of unscheduled DNA synthesis.<br />

Biochemistry 1983;22:5188-5194.<br />

16. Yamamoto H, Uchigata Y, Okamoto H. Streptozotocin and alloxan induce DNA<br />

strand breaks and poly(ADP-ribose) synthetase in pancreatic islets. Nature<br />

1981;294:284-286.<br />

17. Burkart V, Wang ZQ, Radons J, Heller B, Herceg Z, Stingl L, Wagner EF, et al.<br />

Mice lacking the poly(ADP-ribose) polymerase gene are resistant to pancreatic<br />

beta-cell destruction and diabetes development induced by streptozocin. Nat Med<br />

1999;5:314-319.<br />

18. Charron MJ, Bonner-Weir S. Implicating PARP and NAD+ depletion in type I<br />

diabetes. Nat Med 1999;5:269-270.<br />

19. Bergel F, Stock JA. Cyto-active Amino-acid and Peptide Derivatives Part I.<br />

Substituted Phenylalanins. J Chem Soc 1954:2409-2417.<br />

20. Bergel F. Some chemical aspects of abnormal growth. Lect Sci Basis Med<br />

1954;4:54-73.<br />

21. Stock CC. Experimental cancer chemotherapy. Adv Cancer Res 1954;2:425-492.<br />

22. Larionov LF, Shkodinskaja EN, Troosheikina VI, Khokhlov AS, Vasina OS,<br />

Novikova MA. Studies on the anti-tumour activity of p-di-(2-chloroethyl)<br />

aminophenylalanine (sarcolysine). Lancet 1955;269:169-171.<br />

23. Schmidt LH, Fradkin R, Sullivan R, Flowers A. Comparative Pharmacology of<br />

Alkylating Agents. II. Therapeutic Activities of Alkylating Agents and Reference<br />

Compounds against Various Tumor Systems. Cancer Chemother Rep<br />

1965;18:SUPPL 2:403+.<br />

24. Schmidt LH, Fradkin R, Sullivan R, Flowers A. Comparative Pharmacology of<br />

Alkylating Agents. I. Cancer Chemother Rep 1965;18:SUPPL 2:1-401.<br />

25. Schmidt LH, Fradkin R, Sullivan R, Flowers A. Comparative Pharmacology of<br />

Alkylating Agents. 3. Toxicity Data on Monkeys and Dogs. Cancer Chemother<br />

Rep 1965;18:SUPPL 2:1017+.<br />

26. Bosanquet AG. Stability of melphalan solutions during preparation and storage. J<br />

Pharm Sci 1985;74:348-351.<br />

27. Flora KP, Smith SL, Cradock JC. Application of a simple high-performance liquid<br />

chromatographic method for the determination of melphalan in the presence of its<br />

hydrolysis products. J Chromatogr 1979;177:91-97.<br />

28. Goldenberg GJ, Vanstone CL, Israels LG, Iise D, Bihler I. Evidence for a<br />

transport carrier of nitrogen mustard in nitrogen mustard-sensitive and -resistant<br />

L5178Y lymphoblasts. Cancer Res 1970;30:2285-2291.<br />

- 136 -


References<br />

29. Wolpert MK, Ruddon RW. A study on the mechanism of resistance to nitrogen<br />

mustard (HN2) in Ehrlich ascites tumor cells: comparison of uptake of HN2-14-C<br />

into sensitive and resistant cells. Cancer Res 1969;29:873-879.<br />

30. Furner RL, Brown RK. L-phenylalanine mustard (L-PAM): the first 25 years.<br />

Cancer Treat Rep 1980;64:559-574.<br />

31. Hajek R, Adam Z, Vasova I, Kral Z. Evolution of multiple myeloma treatment from<br />

melphalan monotherapy to bone marrow transplantation. Acta Med Austriaca<br />

1996;23:85-91.<br />

32. Alexanian R, Haut A, Khan AU, Lane M, McKelvey EM, Migliore PJ, Stuckey WJ,<br />

Jr., et al. Treatment for multiple myeloma. Combination chemotherapy with<br />

different melphalan dose regimens. Jama 1969;208:1680-1685.<br />

33. Boccadoro M, Marmont F, Tribalto M, Avvisati G, Andriani A, Barbui T, Cantonetti<br />

M, et al. Multiple myeloma: VMCP/VBAP alternating combination chemotherapy<br />

is not superior to melphalan and prednisone even in high-risk patients. J Clin<br />

Oncol 1991;9:444-448.<br />

34. Alexanian R, Barlogie B, Dixon D. High-dose glucocorticoid treatment of resistant<br />

myeloma. Ann Intern Med 1986;105:8-11.<br />

35. Alexanian R, Barlogie B, Tucker S. VAD-based regimens as primary treatment for<br />

multiple myeloma. Am J Hematol 1990;33:86-89.<br />

36. McElwain TJ, Powles RL. High-dose intravenous melphalan for plasma-cell<br />

leukaemia and myeloma. Lancet 1983;2:822-824.<br />

37. Barlogie B, Hall R, Zander A, Dicke K, Alexanian R. High-dose melphalan with<br />

autologous bone marrow transplantation for multiple myeloma. Blood<br />

1986;67:1298-1301.<br />

38. Cunningham D, Paz-Ares L, Gore ME, Malpas J, Hickish T, Nicolson M, Meldrum<br />

M, et al. High-dose melphalan for multiple myeloma: long-term follow-up data. J<br />

Clin Oncol 1994;12:764-768.<br />

39. Harousseau JL, Milpied N, Laporte JP, Collombat P, Facon T, Tigaud JD,<br />

Casassus P, et al. Double-intensive therapy in high-risk multiple myeloma. Blood<br />

1992;79:2827-2833.<br />

40. Lienard D, Eggermont AM, Kroon BB, Schraffordt Koops H, Lejeune FJ. Isolated<br />

limb perfusion in primary and recurrent melanoma: indications and results. Semin<br />

Surg Oncol 1998;14:202-209.<br />

41. Schraffordt Koops H, Eggermont AM, Lienard D, Kroon BB, Hoekstra HJ, Van<br />

Geel AN, Nieweg OE, et al. Hyperthermic isolated limb perfusion for the<br />

treatment of soft tissue sarcomas. Semin Surg Oncol 1998;14:210-214.<br />

42. Weinreich DM, Alexander HR. Transarterial perfusion of liver metastases. Semin<br />

Oncol 2002;29:136-144.<br />

- 137 -


References<br />

43. Ausman RK. Development of a technic for isolated perfusion of the liver. N Y<br />

State J Med 1961;61:3993-3997.<br />

44. Healey JE, Jr., Smith JL, Jr., Clark L, Jr., Stehlin JS, Jr., White EC. Hepatic tissue<br />

tolerance to thio-TEPA administered by the isolation-perfusion technique. J Surg<br />

Res 1961;1:111-116.<br />

45. Breedis C, Young G. The blood supply of neoplasms in the liver. Am J Pathol<br />

1954;30:969-977.<br />

46. Christoforidis D, Martinet O, Lejeune FJ, Mosimann F. Isolated liver perfusion for<br />

non-resectable liver tumours: a review. Eur J Surg Oncol 2002;28:875-890.<br />

47. Kowal CD, Bertino JR. Possible benefits of hyperthermia to chemotherapy.<br />

Cancer Res 1979;39:2285-2289.<br />

48. Giovanella BC, Stehlin JS, Jr., Morgan AC. Selective lethal effect of supranormal<br />

temperatures on human neoplastic cells. Cancer Res 1976;36:3944-3950.<br />

49. van der Veen AH, de Wilt JH, Eggermont AM, van Tiel ST, Seynhaeve AL, ten<br />

Hagen TL. TNF-alpha augments intratumoural concentrations of doxorubicin in<br />

TNF-alpha-based isolated limb perfusion in rat sarcoma models and enhances<br />

anti-tumour effects. Br J Cancer 2000;82:973-980.<br />

50. Marinelli A, van Dierendonck JH, van Brakel GM, Irth H, Kuppen PJ, Tjaden UR,<br />

van de Velde CJ. Increasing the effective concentration of melphalan in<br />

experimental rat liver tumours: comparison of isolated liver perfusion and hepatic<br />

artery infusion. Br J Cancer 1991;64:1069-1075.<br />

51. Marinelli A, van de Velde CJ, Kuppen PJ, Franken HC, Souverijn JH, Eggermont<br />

AM. A comparative study of isolated liver perfusion versus hepatic artery infusion<br />

with mitomycin C in rats. Br J Cancer 1990;62:891-896.<br />

52. Skibba JL, Condon RE. Hyperthermic isolation-perfusion in vivo of the canine<br />

liver. Cancer 1983;51:1303-1309.<br />

53. Skibba JL, Quebbeman EJ. Tumoricidal effects and patient survival after<br />

hyperthermic liver perfusion. Arch Surg 1986;121:1266-1271.<br />

54. Hafstrom LR, Holmberg SB, Naredi PL, Lindner PG, Bengtsson A, Tidebrant G,<br />

Schersten TS. Isolated hyperthermic liver perfusion with chemotherapy for liver<br />

malignancy. Surg Oncol 1994;3:103-108.<br />

55. Hafstrom L, Engaras B, Holmberg SB, Gustavsson B, Jonsson PE, Lindner P,<br />

Naredi P, et al. Treatment of liver metastases from colorectal cancer with hepatic<br />

artery occlusion, intraportal 5-fluorouracil infusion, and oral allopurinol. A<br />

randomized clinical trial. Cancer 1994;74:2749-2756.<br />

56. Marinelli A, de Brauw LM, Beerman H, Keizer HJ, van Bockel JH, Tjaden UR, van<br />

de Velde CJ. Isolated liver perfusion with mitomycin C in the treatment of<br />

colorectal cancer metastases confined to the liver. Jpn J Clin Oncol 1996;26:341-<br />

350.<br />

- 138 -


References<br />

57. Leff RS, Thompson JM, Johnson DB, Mosley KR, Daly MB, Knight WAd, Ruxer<br />

RL, Jr., et al. Phase II trial of high-dose melphalan and autologous bone marrow<br />

transplantation for metastatic colon carcinoma. J Clin Oncol 1986;4:1586-1591.<br />

58. de Vries MR, Rinkes IH, van de Velde CJ, Wiggers T, Tollenaar RA, Kuppen PJ,<br />

Vahrmeijer AL, et al. Isolated hepatic perfusion with tumor necrosis factor alpha<br />

and melphalan: experimental studies in pigs and phase I data from humans.<br />

Recent Results Cancer Res 1998;147:107-119.<br />

59. Alexander HR, Brown CK, Bartlett DL, Libutti SK, Figg WD, Raje S, Turner E.<br />

Augmented capillary leak during isolated hepatic perfusion (IHP) occurs via tumor<br />

necrosis factor-independent mechanisms. Clin Cancer Res 1998;4:2357-2362.<br />

60. Lienard D, Ewalenko P, Delmotte JJ, Renard N, Lejeune FJ. High-dose<br />

recombinant tumor necrosis factor alpha in combination with interferon gamma<br />

and melphalan in isolation perfusion of the limbs for melanoma and sarcoma. J<br />

Clin Oncol 1992;10:52-60.<br />

61. Lienard D, Lejeune FJ, Ewalenko P. In transit metastases of malignant<br />

melanoma treated by high dose rTNF alpha in combination with interferongamma<br />

and melphalan in isolation perfusion. World J Surg 1992;16:234-240.<br />

62. Alexander HR, Jr., Fraker DL, Bartlett DL. Isolated limb perfusion for malignant<br />

melanoma. Semin Surg Oncol 1996;12:416-428.<br />

63. Eggermont AM, Schraffordt Koops H, Klausner JM, Kroon BB, Schlag PM,<br />

Lienard D, van Geel AN, et al. Isolated limb perfusion with tumor necrosis factor<br />

and melphalan for limb salvage in 186 patients with locally advanced soft tissue<br />

extremity sarcomas. The cumulative multicenter European experience. Ann Surg<br />

1996;224:756-764;.<br />

64. Eggermont AM, Schraffordt Koops H, Lienard D, Kroon BB, van Geel AN,<br />

Hoekstra HJ, Lejeune FJ. Isolated limb perfusion with high-dose tumor necrosis<br />

factor-alpha in combination with interferon-gamma and melphalan for<br />

nonresectable extremity soft tissue sarcomas: a multicenter trial. J Clin Oncol<br />

1996;14:2653-2665.<br />

65. Lindner P, Fjalling M, Hafstrom L, Kierulff-Nielsen H, Mattsson J, Schersten T,<br />

Rizell M, et al. Isolated hepatic perfusion with extracorporeal oxygenation using<br />

hyperthermia, tumour necrosis factor alpha and melphalan. Eur J Surg Oncol<br />

1999;25:179-185.<br />

66. Hoption Cann SA, van Netten JP, van Netten C. Dr William Coley and tumour<br />

regression: a place in history or in the future. Postgrad Med J 2003;79:672-680.<br />

67. Nauts HC, Fowler GA, Bogatko FH. A review of the influence of bacterial infection<br />

and of bacterial products (Coley's toxins) on malignant tumors in man; a critical<br />

analysis of 30 inoperable cases treated by Coley's mixed toxins, in which<br />

diagnosis was confirmed by microscopic examination selected for special study.<br />

Acta Med Scand Suppl 1953;144:1-103.<br />

- 139 -


References<br />

68. Carswell EA, Old LJ, Kassel RL, Green S, Fiore N, Williamson B. An endotoxininduced<br />

serum factor that causes necrosis of tumors. Proc Natl Acad Sci U S A<br />

1975;72:3666-3670.<br />

69. Old LJ. Tumor necrosis factor (TNF). Science 1985;230:630-632.<br />

70. Beutler B, Cerami A. The biology of cachectin/TNF--a primary mediator of the<br />

host response. Annu Rev Immunol 1989;7:625-655.<br />

71. Black RA, Rauch CT, Kozlosky CJ, Peschon JJ, Slack JL, Wolfson MF, Castner<br />

BJ, et al. A metalloproteinase disintegrin that releases tumour-necrosis factoralpha<br />

from cells. Nature 1997;385:729-733.<br />

72. Black RA. Tumor necrosis factor-alpha converting enzyme. Int J Biochem Cell<br />

Biol 2002;34:1-5.<br />

73. Decker K. Biologically active products of stimulated liver macrophages (Kupffer<br />

cells). Eur J Biochem 1990;192:245-261.<br />

74. Djeu JY, Serbousek D, Blanchard DK. Release of tumor necrosis factor by<br />

human polymorphonuclear leukocytes. Blood 1990;76:1405-1409.<br />

75. Djeu JY. Tumor necrosis factor and Candida albicans. Behring Inst Mitt<br />

1991:222-227.<br />

76. Djeu JY, Matsushima K, Oppenheim JJ, Shiotsuki K, Blanchard DK. Functional<br />

activation of human neutrophils by recombinant monocyte-derived neutrophil<br />

chemotactic factor/IL-8. J Immunol 1990;144:2205-2210.<br />

77. Dubravec DB, Spriggs DR, Mannick JA, Rodrick ML. Circulating human<br />

peripheral blood granulocytes synthesize and secrete tumor necrosis factor<br />

alpha. Proc Natl Acad Sci U S A 1990;87:6758-6761.<br />

78. Ohno I, Tanno Y, Yamauchi K, Takishima T. Production of tumour necrosis factor<br />

by mastocytoma P815 cells. Immunology 1990;69:312-315.<br />

79. Gordon JR, Galli SJ. Mast cells as a source of both preformed and<br />

immunologically inducible TNF-alpha/cachectin. Nature 1990;346:274-276.<br />

80. Steffen M, Ottmann OG, Moore MA. Simultaneous production of tumor necrosis<br />

factor-alpha and lymphotoxin by normal T cells after induction with IL-2 and anti-<br />

T3. J Immunol 1988;140:2621-2624.<br />

81. Kobayashi Y, Asada M, Osawa T. Production of lymphotoxin and tumour necrosis<br />

factor by a T-cell hybridoma. Immunology 1987;60:213-217.<br />

82. Brennan FM, Maini RN, Feldmann M. TNF alpha--a pivotal role in rheumatoid<br />

arthritis? Br J Rheumatol 1992;31:293-298.<br />

83. Shirai T, Yamaguchi H, Ito H, Todd CW, Wallace RB. Cloning and expression in<br />

Escherichia coli of the gene for human tumour necrosis factor. Nature<br />

1985;131:803-806.<br />

- 140 -


References<br />

84. Pennica D, Nedwin GE, Hayflick JS, Seeburg PH, Derynck R, Palladino MA, Kohr<br />

WJ, et al. Human tumour necrosis factor: precursor structure, expression and<br />

homology to lymphotoxin. Nature 1984;312:724-729.<br />

85. Marmenout A, Fransen L, Tavernier J, Van der Heyden J, Tizard R, Kawashima<br />

E, Shaw A, et al. Molecular cloning and expression of human tumor necrosis<br />

factor and comparison with mouse tumor necrosis factor. Eur J Biochem<br />

1985;152:515-522.<br />

86. Fiers W. Tumor necrosis factor. Characterization at the molecular, cellular and in<br />

vivo level. FEBS Lett 1991;285:199-212.<br />

87. Camussi G, Albano E, Tetta C, Bussolino F. The molecular action of tumor<br />

necrosis factor-alpha. Eur J Biochem 1991;202:3-14.<br />

88. Beutler B, Cerami A. Tumor necrosis, cachexia, shock, and inflammation: a<br />

common mediator. Annu Rev Biochem 1988;57:505-518.<br />

89. Old LJ. Tumour necrosis factor. Polypeptide mediator network. Nature<br />

1987;326:330-331.<br />

90. Cerami A, Beutler B. The role of cachectin/TNF in endotoxic shock and cachexia.<br />

Immunol Today 1988;9:28-31.<br />

91. Lejeune FJ, Ruegg C, Lienard D. Clinical applications of TNF-alpha in cancer.<br />

Curr Opin Immunol 1998;10:573-580.<br />

92. Williams TW, Granger GA. Lymphocyte in vitro cytotoxicity: lymphotoxins of<br />

several mammalian species. Nature 1968;219:1076-1077.<br />

93. Itoh N, Yonehara S, Ishii A, Yonehara M, Mizushima S, Sameshima M, Hase A,<br />

et al. The polypeptide encoded by the cDNA for human cell surface antigen Fas<br />

can mediate apoptosis. Cell 1991;66:233-243.<br />

94. Wiley SR, Schooley K, Smolak PJ, Din WS, Huang CP, Nicholl JK, Sutherland<br />

GR, et al. Identification and characterization of a new member of the TNF family<br />

that induces apoptosis. Immunity 1995;3:673-682.<br />

95. Aggarwal BB. Signalling pathways of the TNF superfamily: a double-edged<br />

sword. Nat Rev Immunol 2003;3:745-756.<br />

96. Gray PW, Aggarwal BB, Benton CV, Bringman TS, Henzel WJ, Jarrett JA, Leung<br />

DW, et al. Cloning and expression of cDNA for human lymphotoxin, a lymphokine<br />

with tumour necrosis activity. Nature 1984;312:721-724.<br />

97. Aggarwal BB, Henzel WJ, Moffat B, Kohr WJ, Harkins RN. Primary structure of<br />

human lymphotoxin derived from 1788 lymphoblastoid cell line. J Biol Chem<br />

1985;260:2334-2344.<br />

98. Aggarwal BB, Moffat B, Harkins RN. Human lymphotoxin. Production by a<br />

lymphoblastoid cell line, purification, and initial characterization. J Biol Chem<br />

1984;259:686-691.<br />

- 141 -


References<br />

99. Locksley RM, Killeen N, Lenardo MJ. The TNF and TNF receptor superfamilies:<br />

integrating mammalian biology. Cell 2001;104:487-501.<br />

100. Aggarwal BB. Comparative analysis of the structure and function of TNF-alpha<br />

and TNF-beta. Immunol Ser 1992;56:61-78.<br />

101. Yonehara S, Ishii A, Yonehara M. A cell-killing monoclonal antibody (anti-Fas) to<br />

a cell surface antigen co-downregulated with the receptor of tumor necrosis<br />

factor. J Exp Med 1989;169:1747-1756.<br />

102. Pitti RM, Marsters SA, Ruppert S, Donahue CJ, Moore A, Ashkenazi A. Induction<br />

of apoptosis by Apo-2 ligand, a new member of the tumor necrosis factor<br />

cytokine family. J Biol Chem 1996;271:12687-12690.<br />

103. Trauth BC, Klas C, Peters AM, Matzku S, Moller P, Falk W, Debatin KM, et al.<br />

Monoclonal antibody-mediated tumor regression by induction of apoptosis.<br />

Science 1989;245:301-305.<br />

104. Oehm A, Behrmann I, Falk W, Pawlita M, Maier G, Klas C, Li-Weber M, et al.<br />

Purification and molecular cloning of the APO-1 cell surface antigen, a member of<br />

the tumor necrosis factor/nerve growth factor receptor superfamily. Sequence<br />

identity with the Fas antigen. J Biol Chem 1992;267:10709-10715.<br />

105. Suda T, Takahashi T, Golstein P, Nagata S. Molecular cloning and expression of<br />

the Fas ligand, a novel member of the tumor necrosis factor family. Cell<br />

1993;75:1169-1178.<br />

106. Takahashi T, Tanaka M, Inazawa J, Abe T, Suda T, Nagata S. Human Fas<br />

ligand: gene structure, chromosomal location and species specificity. Int Immunol<br />

1994;6:1567-1574.<br />

107. Deushi T, Yamaguchi T, Kamiya K, Iwasaki A, Mizoguchi T, Nakayama M,<br />

Watanabe I, et al. A new aminoglycoside antibiotic, sannamycin C and its 4-Nglycyl<br />

derivative. J Antibiot (Tokyo) 1980;33:1274-1280.<br />

108. Galle PR, Hofmann WJ, Walczak H, Schaller H, Otto G, Stremmel W, Krammer<br />

PH, et al. Involvement of the CD95 (APO-1/Fas) receptor and ligand in liver<br />

damage. J Exp Med 1995;182:1223-1230.<br />

109. Jaeschke H. Inflammation in response to hepatocellular apoptosis. Hepatology<br />

2002;35:964-966.<br />

110. Pinkoski MJ, Brunner T, Green DR, Lin T. Fas and Fas ligand in gut and liver. Am<br />

J Physiol Gastrointest Liver Physiol 2000;278:G354-G366.<br />

111. Leist M, Gantner F, Kunstle G, Bohlinger I, Tiegs G, Bluethmann H, Wendel A.<br />

The 55-kD tumor necrosis factor receptor and CD95 independently signal murine<br />

hepatocyte apoptosis and subsequent liver failure. Mol Med 1996;2:109-124.<br />

112. Takehara T, Hayashi N. Fas and fas ligand in human hepatocellular carcinoma. J<br />

Gastroenterol 2001;36:727-728.<br />

- 142 -


References<br />

113. Alderson MR, Armitage RJ, Maraskovsky E, Tough TW, Roux E, Schooley K,<br />

Ramsdell F, et al. Fas transduces activation signals in normal human T<br />

lymphocytes. J Exp Med 1993;178:2231-2235.<br />

114. Desbarats J, Wade T, Wade WF, Newell MK. Dichotomy between naive and<br />

memory CD4(+) T cell responses to Fas engagement. Proc Natl Acad Sci U S A<br />

1999;96:8104-8109.<br />

115. Freiberg RA, Spencer DM, Choate KA, Duh HJ, Schreiber SL, Crabtree GR,<br />

Khavari PA. Fas signal transduction triggers either proliferation or apoptosis in<br />

human fibroblasts. J Invest Dermatol 1997;108:215-219.<br />

116. Desbarats J, Newell MK. Fas engagement accelerates liver regeneration after<br />

partial hepatectomy. Nat Med 2000;6:920-923.<br />

117. Ibuki N, Yamamoto K, Yabushita K, Okano N, Okamoto R, Shimada N, Hakoda<br />

T, et al. In situ expression of Granzyme B and Fas-ligand in the liver of viral<br />

hepatitis. Liver 2002;22:198-204.<br />

118. de Vries EG, Timmer T, Mulder NH, van Geelen CM, van der Graaf WT,<br />

Spierings DC, de Hooge MN, et al. Modulation of death receptor pathways in<br />

oncology. Drugs Today (Barc) 2003;39 Suppl C:95-109.<br />

119. Wang XZ, Chen XC, Chen YX, Zhang LJ, Li D, Chen FL, Chen ZX, et al.<br />

Overexpression of HBxAg in hepatocellular carcinoma and its relationship with<br />

Fas/FasL system. World J Gastroenterol 2003;9:2671-2675.<br />

120. Strand S, Hofmann WJ, Hug H, Muller M, Otto G, Strand D, Mariani SM, et al.<br />

Lymphocyte apoptosis induced by CD95 (APO-1/Fas) ligand-expressing tumor<br />

cells--a mechanism of immune evasion? Nat Med 1996;2:1361-1366.<br />

121. Muller M, Strand S, Hug H, Heinemann EM, Walczak H, Hofmann WJ, Stremmel<br />

W, et al. Drug-induced apoptosis in hepatoma cells is mediated by the CD95<br />

(APO-1/Fas) receptor/ligand system and involves activation of wild-type p53. J<br />

Clin Invest 1997;99:403-413.<br />

122. Kischkel FC, Lawrence DA, Chuntharapai A, Schow P, Kim KJ, Ashkenazi A.<br />

Apo2L/TRAIL-dependent recruitment of endogenous FADD and caspase-8 to<br />

death receptors 4 and 5. Immunity 2000;12:611-620.<br />

123. Mildner M, Eckhart L, Lengauer B, Tschachler E. Hepatocyte growth<br />

factor/scatter factor inhibits UVB-induced apoptosis of human keratinocytes but<br />

not of keratinocyte-derived cell lines via the phosphatidylinositol 3-kinase/AKT<br />

pathway. J Biol Chem 2002;277:14146-14152.<br />

124. Arbuckle E, Langlois NE, Eremin O, Heys SD. Evidence for Fas counter attack in<br />

vivo from a study of colorectal cancer. Oncol Rep 2000;7:45-47.<br />

125. O'Connell J, O'Sullivan GC, Collins JK, Shanahan F. The Fas counterattack: Fasmediated<br />

T cell killing by colon cancer cells expressing Fas ligand. J Exp Med<br />

1996;184:1075-1082.<br />

- 143 -


References<br />

126. Cardi G, Heaney JA, Schned AR, Ernstoff MS. Expression of Fas(APO-1/CD95)<br />

in tumor-infiltrating and peripheral blood lymphocytes in patients with renal cell<br />

carcinoma. Cancer Res 1998;58:2078-2080.<br />

127. Hahne M, Rimoldi D, Schroter M, Romero P, Schreier M, French LE, Schneider<br />

P, et al. Melanoma cell expression of Fas(Apo-1/CD95) ligand: implications for<br />

tumor immune escape. Science 1996;274:1363-1366.<br />

128. Pan G, O'Rourke K, Chinnaiyan AM, Gentz R, Ebner R, Ni J, Dixit VM. The<br />

receptor for the cytotoxic ligand TRAIL. Science 1997;276:111-113.<br />

129. Sheridan JP, Marsters SA, Pitti RM, Gurney A, Skubatch M, Baldwin D,<br />

Ramakrishnan L, et al. Control of TRAIL-induced apoptosis by a family of<br />

signaling and decoy receptors. Science 1997;277:818-821.<br />

130. Walczak H, Degli-Esposti MA, Johnson RS, Smolak PJ, Waugh JY, Boiani N,<br />

Timour MS, et al. TRAIL-R2: a novel apoptosis-mediating receptor for TRAIL.<br />

Embo J 1997;16:5386-5397.<br />

131. Degli-Esposti MA, Smolak PJ, Walczak H, Waugh J, Huang CP, DuBose RF,<br />

Goodwin RG, et al. Cloning and characterization of TRAIL-R3, a novel member of<br />

the emerging TRAIL receptor family. J Exp Med 1997;186:1165-1170.<br />

132. Degli-Esposti MA, Dougall WC, Smolak PJ, Waugh JY, Smith CA, Goodwin RG.<br />

The novel receptor TRAIL-R4 induces NF-kappaB and protects against TRAILmediated<br />

apoptosis, yet retains an incomplete death domain. Immunity<br />

1997;7:813-820.<br />

133. Marsters SA, Sheridan JP, Pitti RM, Huang A, Skubatch M, Baldwin D, Yuan J, et<br />

al. A novel receptor for Apo2L/TRAIL contains a truncated death domain. Curr<br />

Biol 1997;7:1003-1006.<br />

134. Golstein P. Cell death: TRAIL and its receptors. Curr Biol 1997;7:R750-R753.<br />

135. Griffith TS, Lynch DH. TRAIL: a molecule with multiple receptors and control<br />

mechanisms. Curr Opin Immunol 1998;10:559-563.<br />

136. Sprick MR, Weigand MA, Rieser E, Rauch CT, Juo P, Blenis J, Krammer PH, et<br />

al. FADD/MORT1 and caspase-8 are recruited to TRAIL receptors 1 and 2 and<br />

are essential for apoptosis mediated by TRAIL receptor 2. Immunity 2000;12:599-<br />

609.<br />

137. Kuang AA, Diehl GE, Zhang J, Winoto A. FADD is required for DR4- and DR5mediated<br />

apoptosis: lack of trail-induced apoptosis in FADD-deficient mouse<br />

embryonic fibroblasts. J Biol Chem 2000;275:25065-25068.<br />

138. Peter ME. The TRAIL DISCussion: It is FADD and caspase-8! Cell Death Differ<br />

2000;7:759-760.<br />

139. Holler N, Zaru R, Micheau O, Thome M, Attinger A, Valitutti S, Bodmer JL, et al.<br />

Fas triggers an alternative, caspase-8-independent cell death pathway using the<br />

kinase RIP as effector molecule. Nat Immunol 2000;1:489-495.<br />

- 144 -


References<br />

140. Walczak H, Miller RE, Ariail K, Gliniak B, Griffith TS, Kubin M, Chin W, et al.<br />

Tumoricidal activity of tumor necrosis factor-related apoptosis-inducing ligand in<br />

vivo. Nat Med 1999;5:157-163.<br />

141. Tanaka S, Sugimachi K, Shirabe K, Shimada M, Wands JR. Expression and<br />

antitumor effects of TRAIL in human cholangiocarcinoma. Hepatology<br />

2000;32:523-527.<br />

142. Yamanaka T, Shiraki K, Sugimoto K, Ito T, Fujikawa K, Ito M, Takase K, et al.<br />

Chemotherapeutic agents augment TRAIL-induced apoptosis in human<br />

hepatocellular carcinoma cell lines. Hepatology 2000;32:482-490.<br />

143. Gores GJ, Kaufmann SH. Is TRAIL hepatotoxic? Hepatology 2001;34:3-6.<br />

144. Mori E, Thomas M, Motoki K, Nakazawa K, Tahara T, Tomizuka K, Ishida I, et al.<br />

Human normal hepatocytes are susceptible to apoptosis signal mediated by both<br />

TRAIL-R1 and TRAIL-R2. Cell Death Differ 2004;11:203-207.<br />

145. Zheng SJ, Wang P, Tsabary G, Chen YH. Critical roles of TRAIL in hepatic cell<br />

death and hepatic inflammation. J Clin Invest 2004;113:58-64.<br />

146. Gonzalez VM, Fuertes MA, Alonso C, Perez JM. Is cisplatin-induced cell death<br />

always produced by apoptosis? Mol Pharmacol 2001;59:657-663.<br />

147. Oh KW, Qian T, Brenner DA, Lemasters JJ. Salicylate enhances necrosis and<br />

apoptosis mediated by the mitochondrial permeability transition. Toxicol Sci<br />

2003;73:44-52.<br />

148. Mundt B, Kuhnel F, Zender L, Paul Y, Tillmann H, Trautwein C, Manns MP, et al.<br />

Involvement of TRAIL and its receptors in viral hepatitis. Faseb J 2003;17:94-96.<br />

149. Higuchi H, Bronk SF, Taniai M, Canbay A, Gores GJ. Cholestasis Increases<br />

Tumor Necrosis Factor-Related Apoptotis-Inducing Ligand (TRAIL)-R2/DR5<br />

Expression and Sensitizes the Liver to TRAIL- Mediated Cytotoxicity. J<br />

Pharmacol Exp Ther 2002;303:461-467.<br />

150. Lin T, Gu J, Zhang L, Huang X, Stephens LC, Curley SA, Fang B. Targeted<br />

expression of green fluorescent protein/tumor necrosis factor-related apoptosisinducing<br />

ligand fusion protein from human telomerase reverse transcriptase<br />

promoter elicits antitumor activity without toxic effects on primary human<br />

hepatocytes. Cancer Res 2002;62:3620-3625.<br />

151. Srivastava RK. TRAIL/Apo-2L: mechanisms and clinical applications in cancer.<br />

Neoplasia 2001;3:535-546.<br />

152. Nagane M, Huang HJ, Cavenee WK. The potential of TRAIL for cancer<br />

chemotherapy. Apoptosis 2001;6:191-197.<br />

153. Kagawa S, He C, Gu J, Koch P, Rha SJ, Roth JA, Curley SA, et al. Antitumor<br />

activity and bystander effects of the tumor necrosis factor-related apoptosisinducing<br />

ligand (TRAIL) gene. Cancer Res 2001;61:3330-3338.<br />

- 145 -


References<br />

154. Lawrence D, Shahrokh Z, Marsters S, Achilles K, Shih D, Mounho B, Hillan K, et<br />

al. Differential hepatocyte toxicity of recombinant Apo2L/TRAIL versions. Nat<br />

Med 2001;7:383-385.<br />

155. Ozoren N, Kim K, Burns TF, Dicker DT, Moscioni AD, El-Deiry WS. The caspase<br />

9 inhibitor Z-LEHD-FMK protects human liver cells while permitting death of<br />

cancer cells exposed to tumor necrosis factor-related apoptosis-inducing ligand.<br />

Cancer Res 2000;60:6259-6265.<br />

156. Hugh TJ, Kinsella AR, Poston GJ. Management strategies for colorectal liver<br />

metastases--Part I. Surg Oncol 1997;6:19-30.<br />

157. Weiss L, Grundmann E, Torhorst J, Hartveit F, Moberg I, Eder M, Fenoglio-<br />

Preiser CM, et al. Haematogenous metastatic patterns in colonic carcinoma: an<br />

analysis of 1541 necropsies. J Pathol 1986;150:195-203.<br />

158. Yoon SS, Tanabe KK. Surgical treatment and other regional treatments for<br />

colorectal cancer liver metastases. Oncologist 1999;4:197-208.<br />

159. Habib NA, Hodgson HJ, Lemoine N, Pignatelli M. A phase I/II study of hepatic<br />

artery infusion with wtp53-CMV-Ad in metastatic malignant liver tumours. Hum<br />

Gene Ther 1999;10:2019-2034.<br />

160. Greenway B. Hepatic metastases from colorectal cancer: resection or not. Br J<br />

Surg 1988;75:513-519.<br />

161. Cady B, Stone MD. The role of surgical resection of liver metastases in colorectal<br />

carcinoma. Semin Oncol 1991;18:399-406.<br />

162. Wagner JS, Adson MA, Van Heerden JA, Adson MH, Ilstrup DM. The natural<br />

history of hepatic metastases from colorectal cancer. A comparison with resective<br />

treatment. Ann Surg 1984;199:502-508.<br />

163. Oxley EM, Ellis H. Prognosis of carcinoma of the large bowel in the presence of<br />

liver metastases. Br J Surg 1969;56:149-152.<br />

164. Flanagan L, Jr., Foster JH. Hepatic resection for metastatic cancer. Am J Surg<br />

1967;113:551-557.<br />

165. Ragnhammar P, Hafstrom L, Nygren P, Glimelius B. A systematic overview of<br />

chemotherapy effects in colorectal cancer. Acta Oncol 2001;40:282-308.<br />

166. Hafstrom L, Holmberg SB, Lindner P, Naredi P, Asztely M, Bengtsson A, Nielsen<br />

HK, et al. Isolated regional liver perfusion with tumour necrosis factor-alpha<br />

followed by hyperthermic melphalan perfusion. Reg Cancer Treat. 1994;7:172-<br />

176.<br />

167. Ahmed N, Berridge MV. Ceramides that mediate apoptosis reduce glucose<br />

uptake and transporter affinity for glucose in human leukaemic cell lines but not in<br />

neutrophils. Pharmacol Toxicol 2000;86:114-121.<br />

- 146 -


References<br />

168. Borel Rinkes IH, de Vries MR, Jonker AM, Swaak TJ, Hack CE, Nooyen PT,<br />

Wiggers T, et al. Isolated hepatic perfusion in the pig with TNF-alpha with and<br />

without melphalan. Br J Cancer 1997;75:1447-1453.<br />

169. Goto A, Shomori K, Ohkumo T, Tanaka F, Sato K, Ito H. Hyperthermia-induced<br />

apoptosis occurs both in a p53 gene-dependent and -independent manner in<br />

three human gastric carcinoma cell lines. Oncol Rep 1999;6:335-339.<br />

170. Kobayashi D, Watanabe N, Yamauchi N, Okamoto T, Tsuji N, Sasaki H, Sato T,<br />

et al. Heat-induced apoptosis via caspase-3 activation in tumour cells carrying<br />

mutant p53. Int J Hyperthermia 2000;16:471-480.<br />

171. van Bree C, van der Maat B, Ceha HM, Franken NA, Haveman J, Bakker PJ.<br />

Inactivation of p53 and of pRb protects human colorectal carcinoma cells against<br />

hyperthermia-induced cytotoxicity and apoptosis. J Cancer Res Clin Oncol<br />

1999;125:549-555.<br />

172. Hafstrom L, Naredi P. Isolated hepatic perfusion with extracorporeal oxygenation<br />

using hyperthermia TNF alpha and melphalan: Swedish experience. Recent<br />

Results Cancer Res 1998;147:120-126.<br />

173. Oldhafer KJ, Lang H, Frerker M, Moreno L, Chavan A, Flemming P, Nadalin S, et<br />

al. First experience and technical aspects of isolated liver perfusion for extensive<br />

liver metastasis. Surgery 1998;123:622-631.<br />

174. Alexander HR, Jr., Bartlett DL, Libutti SK, Fraker DL, Moser T, Rosenberg SA.<br />

Isolated hepatic perfusion with tumor necrosis factor and melphalan for<br />

unresectable cancers confined to the liver. J Clin Oncol 1998;16:1479-1489.<br />

175. Plaat BE, Molenaar WM, Mastik MF, Koudstaal J, van den Berg E, Koops HS,<br />

Hoekstra HJ. Hyperthermic isolated limb perfusion with tumor necrosis factoralpha<br />

and melphalan in patients with locally advanced soft tissue sarcomas:<br />

treatment response and clinical outcome related to changes in proliferation and<br />

apoptosis. Clin Cancer Res 1999;5:1650-1657.<br />

176. Skapek SX, VanDellen AF, McMahon DP, Postels DG, Griffith OW, Bigner DD,<br />

Friedman HS. Melphalan-induced toxicity in nude mice following pretreatment<br />

with buthionine sulfoximine. Cancer Chemother Pharmacol 1991;28:15-21.<br />

177. Laskowitz DT, Elion GB, Dewhirst MW, Griffith OW, Cattley RC, Bigner DD,<br />

Friedman HS. Enhancement of melphalan-induced gastrointestinal toxicity in<br />

mice treated with regional hyperthermia and BSO-mediated glutathione depletion.<br />

Int J Hyperthermia 1992;8:111-120.<br />

178. Smith AC, Liao JT, Page JG, Wientjes MG, Grieshaber CK. Pharmacokinetics of<br />

buthionine sulfoximine (NSC 326231) and its effect on melphalan-induced toxicity<br />

in mice. Cancer Res 1989;49:5385-5391.<br />

179. Kramer RA, Schuller HM, Smith AC, Boyd MR. Effects of buthionine sulfoximine<br />

on the nephrotoxicity of 1-(2-chloroethyl)-3-(trans-4-methylcyclohexyl)-1nitrosourea<br />

(MeCCNU). J Pharmacol Exp Ther 1985;234:498-506.<br />

- 147 -


References<br />

180. Vahrmeijer AL, van Dierendonck JH, Schutrups J, van de Velde CJ, Mulder GJ.<br />

Effect of glutathione depletion on inhibition of cell cycle progression and induction<br />

of apoptosis by melphalan (L-phenylalanine mustard) in human colorectal cancer<br />

cells. Biochem Pharmacol 1999;58:655-664.<br />

181. Guenthner TM, Whalen R, Jevtovic-Todorovic V. Direct measurement of<br />

melphalan conjugation with glutathione: studies with human melanoma cells and<br />

mammalian liver. J Pharmacol Exp Ther 1992;260:1331-1336.<br />

182. Vahrmeijer AL, Snel CA, Steenvoorden DP, Beijnen JH, Pang KS, Schutrups J,<br />

Tirona R, et al. Lack of glutathione conjugation of melphalan in the isolated in situ<br />

liver perfusion in humans. Cancer Res 1996;56:4709-4714.<br />

183. Virchow R. Cellular Pathology as based upon Physiological and Pathological<br />

Histology. Dover Publications, New York 1991:356-382.<br />

184. Glücksmann A. Cell death in normal vertebrate ontogeny. Biol Rev 1951;26:59-<br />

86.<br />

185. Ellis RE, Yuan JY, Horvitz HR. Mechanisms and functions of cell death. Annu<br />

Rev Cell Biol 1991;7:663-698.<br />

186. Saunders JW, Jr. Death in embryonic systems. Science 1966;154:604-612.<br />

187. Kerr JF, Wyllie AH, Currie AR. Apoptosis: a basic biological phenomenon with<br />

wide-ranging implications in tissue kinetics. Br J Cancer 1972;26:239-257.<br />

188. Patel T, Gores GJ. Apoptosis and hepatobiliary disease. Hepatology<br />

1995;21:1725-1741.<br />

189. Majno G, Joris I. Apoptosis, oncosis, and necrosis. An overview of cell death. Am<br />

J Pathol 1995;146:3-15.<br />

190. Schwartzman RA, Cidlowski JA. Apoptosis: the biochemistry and molecular<br />

biology of programmed cell death. Endocr Rev 1993;14:133-151.<br />

191. Thompson CB. Apoptosis in the pathogenesis and treatment of disease. Science<br />

1995;267:1456-1462.<br />

192. Chen G, Goeddel DV. TNF-R1 signaling: a beautiful pathway. Science<br />

2002;296:1634-1635.<br />

193. Krammer PH. CD95(APO-1/Fas)-mediated apoptosis: live and let die. Adv<br />

Immunol 1999;71:163-210.<br />

194. Wajant H. The Fas signaling pathway: more than a paradigm. Science<br />

2002;296:16358-11636.<br />

195. Danial NN, Korsmeyer SJ. Cell death: critical control points. Cell 2004;116:205-<br />

219.<br />

- 148 -


References<br />

196. Earnshaw WC, Martins LM, Kaufmann SH. Mammalian caspases: structure,<br />

activation, substrates, and functions during apoptosis. Annu Rev Biochem<br />

1999;68:383-424.<br />

197. Siegel RM, Muppidi J, Roberts M, Porter M, Wu Z. Death receptor signaling and<br />

autoimmunity. Immunol Res 2003;27:499-512.<br />

198. Fausto N. Liver regeneration. J Hepatol 2000;32:19-31.<br />

199. Hengartner MO. The biochemistry of apoptosis. Nature 2000;407:770-776.<br />

200. Cohen GM. Caspases: the executioners of apoptosis. Biochem J 1997;326:1-16.<br />

201. Galle PR. Apoptosis in liver disease. J Hepatol 1997;27:405-412.<br />

202. Fleisher TA. Apoptosis. Ann Allergy Asthma Immunol 1997;78:245-249.<br />

203. Hall AG. Review: The role of glutathione in the regulation of apoptosis. Eur J Clin<br />

Invest 1999;29:238-245.<br />

204. Wang TH, Wang HS. Apoptosis: (1). Overview and clinical significance. J Formos<br />

Med Assoc 1999;98:381-393.<br />

205. Wang TH, Wang HS. Apoptosis: (2) characteristics of apoptosis. J Formos Med<br />

Assoc 1999;98:531-542.<br />

206. Williams G. SC, Grimes E., McCarthy N. Apoptosis: final control point in cell<br />

biology. Trends in cell biology 1992;2:263-267.<br />

207. Chao DT, Korsmeyer SJ. BCL-2 family: regulators of cell death. Annu Rev<br />

Immunol 1998;16:395-419.<br />

208. Pettmann B, Henderson CE. Neuronal cell death. Neuron 1998;20:633-647.<br />

209. Green DR. Apoptotic pathways: the roads to ruin. Cell 1998;94:695-698.<br />

210. Cohen JJ. Apoptosis. Immunol Today 1993;14:126-130.<br />

211. Leong KG, Karsan A. Signaling pathways mediated by tumor necrosis factor<br />

alpha. Histol Histopathol 2000;15:1303-1325.<br />

212. Leist M, Gantner F, Jilg S, Wendel A. Activation of the 55 kDa TNF receptor is<br />

necessary and sufficient for TNF-induced liver failure, hepatocyte apoptosis, and<br />

nitrite release. J Immunol 1995;154:1307-1316.<br />

213. Laster SM, Wood JG, Gooding LR. Tumor necrosis factor can induce both<br />

apoptic and necrotic forms of cell lysis. J Immunol 1988;141:2629-2634.<br />

214. Itoh N, Nagata S. A novel protein domain required for apoptosis. Mutational<br />

analysis of human Fas antigen. J Biol Chem 1993;268:10932-10937.<br />

215. Natoli G, Costanzo A, Guido F, Moretti F, Levrero M. Apoptotic, non-apoptotic,<br />

and anti-apoptotic pathways of tumor necrosis factor signalling. Biochem<br />

Pharmacol 1998;56:915-920.<br />

- 149 -


References<br />

216. Sidoti-de Fraisse C, Rincheval V, Risler Y, Mignotte B, Vayssiere JL. TNF-alpha<br />

activates at least two apoptotic signaling cascades. Oncogene 1998;17:1639-<br />

1651.<br />

217. Tartaglia LA, Rothe M, Hu YF, Goeddel DV. Tumor necrosis factor's cytotoxic<br />

activity is signaled by the p55 TNF receptor. Cell 1993;73:213-216.<br />

218. Tartaglia LA, Ayres TM, Wong GH, Goeddel DV. A novel domain within the 55 kd<br />

TNF receptor signals cell death. Cell 1993;74:845-853.<br />

219. Jiang Y, Woronicz JD, Liu W, Goeddel DV. Prevention of constitutive TNF<br />

receptor 1 signaling by silencer of death domains. Science 1999;283:543-546.<br />

220. Tschopp J, Martinon F, Hofmann K. Apoptosis: Silencing the death receptors.<br />

Curr Biol 1999;9:R381-384.<br />

221. Hsu H, Xiong J, Goeddel DV. The TNF receptor 1-associated protein TRADD<br />

signals cell death and NF-kappa B activation. Cell 1995;81:495-504.<br />

222. Hsu H, Shu HB, Pan MG, Goeddel DV. TRADD-TRAF2 and TRADD-FADD<br />

interactions define two distinct TNF receptor 1 signal transduction pathways. Cell<br />

1996;84:299-308.<br />

223. Boldin MP, Varfolomeev EE, Pancer Z, Mett IL, Camonis JH, Wallach D. A novel<br />

protein that interacts with the death domain of Fas/APO1 contains a sequence<br />

motif related to the death domain. J Biol Chem 1995;270:7795-7798.<br />

224. Chinnaiyan AM, O'Rourke K, Tewari M, Dixit VM. FADD, a novel death domaincontaining<br />

protein, interacts with the death domain of Fas and initiates apoptosis.<br />

Cell 1995;81:505-512.<br />

225. Muzio M, Chinnaiyan AM, Kischkel FC, O'Rourke K, Shevchenko A, Ni J, Scaffidi<br />

C, et al. FLICE, a novel FADD-homologous ICE/CED-3-like protease, is recruited<br />

to the CD95 (Fas/APO-1) death--inducing signaling complex. Cell 1996;85:817-<br />

827.<br />

226. Hofmann K, Bucher P, Tschopp J. The CARD domain: a new apoptotic signalling<br />

motif. Trends Biochem Sci 1997;22:155-156.<br />

227. Medema JP, Scaffidi C, Kischkel FC, Shevchenko A, Mann M, Krammer PH,<br />

Peter ME. FLICE is activated by association with the CD95 death-inducing<br />

signaling complex (DISC). Embo J 1997;16:2794-2804.<br />

228. Wajant H, Johannes FJ, Haas E, Siemienski K, Schwenzer R, Schubert G, Weiss<br />

T, et al. Dominant-negative FADD inhibits TNFR60-, Fas/Apo1- and TRAIL-<br />

R/Apo2-mediated cell death but not gene induction. Curr Biol 1998;8:113-116.<br />

229. Muzio M, Stockwell BR, Stennicke HR, Salvesen GS, Dixit VM. An induced<br />

proximity model for caspase-8 activation. J Biol Chem 1998;273:2926-2930.<br />

230. Rath PC, Aggarwal BB. TNF-induced signaling in apoptosis. J Clin Immunol<br />

1999;19:350-364.<br />

- 150 -


References<br />

231. Pan G, O'Rourke K, Dixit VM. Caspase-9, Bcl-XL, and Apaf-1 form a ternary<br />

complex. J Biol Chem 1998;273:5841-5845.<br />

232. Srinivasula SM, Hegde R, Saleh A, Datta P, Shiozaki E, Chai J, Lee RA, et al. A<br />

conserved XIAP-interaction motif in caspase-9 and Smac/DIABLO regulates<br />

caspase activity and apoptosis. Nature 2001;410:112-116.<br />

233. Shi Y. A structural view of mitochondria-mediated apoptosis. Nat Struct Biol<br />

2001;8:394-401.<br />

234. Joza N, Susin SA, Daugas E, Stanford WL, Cho SK, Li CY, Sasaki T, et al.<br />

Essential role of the mitochondrial apoptosis-inducing factor in programmed cell<br />

death. Nature 2001;410:549-554.<br />

235. Harris MH, Thompson CB. The role of the Bcl-2 family in the regulation of outer<br />

mitochondrial membrane permeability. Cell Death Differ 2000;7:1182-1191.<br />

236. Li H, Zhu H, Xu CJ, Yuan J. Cleavage of BID by caspase 8 mediates the<br />

mitochondrial damage in the Fas pathway of apoptosis. Cell 1998;94:491-501.<br />

237. Zou H, Henzel WJ, Liu X, Lutschg A, Wang X. Apaf-1, a human protein<br />

homologous to C. elegans CED-4, participates in cytochrome c-dependent<br />

activation of caspase-3. Cell 1997;90:405-413.<br />

238. Srinivasula SM, Ahmad M, Fernandes-Alnemri T, Alnemri ES. Autoactivation of<br />

procaspase-9 by Apaf-1-mediated oligomerization. Mol Cell 1998;1:949-957.<br />

239. Li P, Nijhawan D, Budihardjo I, Srinivasula SM, Ahmad M, Alnemri ES, Wang X.<br />

Cytochrome c and dATP-dependent formation of Apaf-1/caspase-9 complex<br />

initiates an apoptotic protease cascade. Cell 1997;91:479-489.<br />

240. Budihardjo I, Oliver H, Lutter M, Luo X, Wang X. Biochemical pathways of<br />

caspase activation during apoptosis. Annu Rev Cell Dev Biol 1999;15:269-290.<br />

241. Lorenzo HK, Susin SA, Penninger J, Kroemer G. Apoptosis inducing factor (AIF):<br />

a phylogenetically old, caspase-independent effector of cell death. Cell Death<br />

Differ 1999;6:516-524.<br />

242. Bigda J, Beletsky I, Brakebusch C, Varfolomeev Y, Engelmann H, Holtmann H,<br />

Wallach D. Dual role of the p75 tumor necrosis factor (TNF) receptor in TNF<br />

cytotoxicity. J Exp Med 1994;180:445-460.<br />

243. Medvedev AE, Sundan A, Espevik T. Involvement of the tumor necrosis factor<br />

receptor p75 in mediating cytotoxicity and gene regulating activities. Eur J<br />

Immunol 1994;24:2842-2849.<br />

244. Heller RA, Song K, Fan N, Chang DJ. The p70 tumor necrosis factor receptor<br />

mediates cytotoxicity. Cell 1992;70:47-56.<br />

245. Lucas R, Garcia I, Donati YR, Hribar M, Mandriota SJ, Giroud C, Buurman WA,<br />

et al. Both TNF receptors are required for direct TNF-mediated cytotoxicity in<br />

microvascular endothelial cells. Eur J Immunol 1998;28:3577-3586.<br />

- 151 -


References<br />

246. Zheng L, Fisher G, Miller RE, Peschon J, Lynch DH, Lenardo MJ. Induction of<br />

apoptosis in mature T cells by tumour necrosis factor. Nature 1995;377:348-351.<br />

247. Shalaby MR, Sundan A, Loetscher H, Brockhaus M, Lesslauer W, Espevik T.<br />

Binding and regulation of cellular functions by monoclonal antibodies against<br />

human tumor necrosis factor receptors. J Exp Med 1990;172:1517-1520.<br />

248. Küsters S, Tiegs G, Alexopoulou L, Pasparakis M, Douni E, Kunstle G,<br />

Bluethmann H, et al. In vivo evidence for a functional role of both tumor necrosis<br />

factor (TNF) receptors and transmembrane TNF in experimental hepatitis. Eur J<br />

Immunol 1997;27:2870-2875.<br />

249. Künstle G, Hentze H, Germann PG, Tiegs G, Meergans T, Wendel A.<br />

Concanavalin A hepatotoxicity in mice: tumor necrosis factor-mediated organ<br />

failure independent of caspase-3-like protease activation. Hepatology<br />

1999;30:1241-1251.<br />

250. Collawn JF, Stangel M, Kuhn LA, Esekogwu V, Jing SQ, Trowbridge IS, Tainer<br />

JA. Transferrin receptor internalization sequence YXRF implicates a tight turn as<br />

the structural recognition motif for endocytosis. Cell 1990;63:1061-1072.<br />

251. Espevik T, Brockhaus M, Loetscher H, Nonstad U, Shalaby R. Characterization<br />

of binding and biological effects of monoclonal antibodies against a human tumor<br />

necrosis factor receptor. J Exp Med 1990;171:415-426.<br />

252. Barbara JA, Smith WB, Gamble JR, Van Ostade X, Vandenabeele P, Tavernier<br />

J, Fiers W, et al. Dissociation of TNF-alpha cytotoxic and proinflammatory<br />

activities by p55 receptor- and p75 receptor-selective TNF-alpha mutants. Embo<br />

J 1994;13:843-850.<br />

253. Van Ostade X, Vandenabeele P, Tavernier J, Fiers W. Human tumor necrosis<br />

factor mutants with preferential binding to and activity on either the R55 or R75<br />

receptor. Eur J Biochem 1994;220:771-779.<br />

254. Loetscher H, Stueber D, Banner D, Mackay F, Lesslauer W. Human tumor<br />

necrosis factor alpha (TNF alpha) mutants with exclusive specificity for the 55kDa<br />

or 75-kDa TNF receptors. J Biol Chem 1993;268:26350-26357.<br />

255. Tartaglia LA, Pennica D, Goeddel DV. Ligand passing: the 75-kDa tumor<br />

necrosis factor (TNF) receptor recruits TNF for signaling by the 55-kDa TNF<br />

receptor. J Biol Chem 1993;268:18542-18548.<br />

256. Vandenabeele P, Declercq W, Beyaert R, Fiers W. Two tumour necrosis factor<br />

receptors: structure and function. Trends Cell Biol 1995;5:392-399.<br />

257. Haridas V, Darnay BG, Natarajan K, Heller R, Aggarwal BB. Overexpression of<br />

the p80 TNF receptor leads to TNF-dependent apoptosis, nuclear factor-kappa B<br />

activation, and c-Jun kinase activation. J Immunol 1998;160:3152-3162.<br />

258. Declercq W, Denecker G, Fiers W, Vandenabeele P. Cooperation of both TNF<br />

receptors in inducing apoptosis: involvement of the TNF receptor-associated<br />

factor binding domain of the TNF receptor 75. J Immunol 1998;161:390-399.<br />

- 152 -


References<br />

259. Weiss T, Grell M, Siemienski K, Muhlenbeck F, Durkop H, Pfizenmaier K,<br />

Scheurich P, et al. TNFR80-dependent enhancement of TNFR60-induced cell<br />

death is mediated by TNFR-associated factor 2 and is specific for TNFR60. J<br />

Immunol 1998;161:3136-3142.<br />

260. Li X, Yang Y, Ashwell JD. TNF-RII and c-IAP1 mediate ubiquitination and<br />

degradation of TRAF2. Nature 2002;416:345-347.<br />

261. Fotin-Mleczek M, Henkler F, Samel D, Reichwein M, Hausser A, Parmryd I,<br />

Scheurich P, et al. Apoptotic crosstalk of TNF receptors: TNF-R2-induces<br />

depletion of TRAF2 and IAP proteins and accelerates TNF-R1-dependent<br />

activation of caspase-8. J Cell Sci 2002;115:2757-2770.<br />

262. Rothe M, Wong SC, Henzel WJ, Goeddel DV. A novel family of putative signal<br />

transducers associated with the cytoplasmic domain of the 75 kDa tumor<br />

necrosis factor receptor. Cell 1994;78:681-692.<br />

263. Laegreid A, Medvedev A, Nonstad U, Bombara MP, Ranges G, Sundan A,<br />

Espevik T. Tumor necrosis factor receptor p75 mediates cell-specific activation of<br />

nuclear factor kappa B and induction of human cytomegalovirus enhancer. J Biol<br />

Chem 1994;269:7785-7791.<br />

264. Lin Y, Devin A, Rodriguez Y, Liu ZG. Cleavage of the death domain kinase RIP<br />

by caspase-8 prompts TNF-induced apoptosis. Genes Dev 1999;13:2514-2526.<br />

265. Song HY, Regnier CH, Kirschning CJ, Goeddel DV, Rothe M. Tumor necrosis<br />

factor (TNF)-mediated kinase cascades: bifurcation of nuclear factor-kappaB and<br />

c-jun N-terminal kinase (JNK/SAPK) pathways at TNF receptor-associated factor<br />

2. Proc Natl Acad Sci U S A 1997;94:9792-9796.<br />

266. Malinin NL, Boldin MP, Kovalenko AV, Wallach D. MAP3K-related kinase<br />

involved in NF-kappaB induction by TNF, CD95 and IL-1. Nature 1997;385:540-<br />

544.<br />

267. Woronicz JD, Gao X, Cao Z, Rothe M, Goeddel DV. IkappaB kinase-beta: NFkappaB<br />

activation and complex formation with IkappaB kinase-alpha and NIK.<br />

Science 1997;278:866-869.<br />

268. DiDonato JA, Hayakawa M, Rothwarf DM, Zandi E, Karin M. A cytokineresponsive<br />

IkappaB kinase that activates the transcription factor NF-kappaB.<br />

Nature 1997;388:548-554.<br />

269. Mercurio F, Zhu H, Murray BW, Shevchenko A, Bennett BL, Li J, Young DB, et al.<br />

IKK-1 and IKK-2: cytokine-activated IkappaB kinases essential for NF-kappaB<br />

activation. Science 1997;278:860-866.<br />

270. Regnier CH, Song HY, Gao X, Goeddel DV, Cao Z, Rothe M. Identification and<br />

characterization of an IkappaB kinase. Cell 1997;90:373-383.<br />

271. Baldwin AS, Jr. The NF-kappa B and I kappa B proteins: new discoveries and<br />

insights. Annu Rev Immunol 1996;14:649-683.<br />

- 153 -


References<br />

272. Karin M, Ben-Neriah Y. Phosphorylation meets ubiquitination: the control of NF-<br />

[kappa]B activity. Annu Rev Immunol 2000;18:621-663.<br />

273. Ghosh S, Karin M. Missing pieces in the NF-kappaB puzzle. Cell 2002;109<br />

Suppl:S81-96.<br />

274. Vig E, Green M, Liu Y, Donner DB, Mukaida N, Goebl MG, Harrington MA.<br />

Modulation of tumor necrosis factor and interleukin-1-dependent NF-kappaB<br />

activity by mPLK/IRAK. J Biol Chem 1999;274:13077-13084.<br />

275. Feinstein E, Kimchi A, Wallach D, Boldin M, Varfolomeev E. The death domain: a<br />

module shared by proteins with diverse cellular functions. Trends Biochem Sci<br />

1995;20:342-344.<br />

276. Trofimova M, Sprenkle AB, Green M, Sturgill TW, Goebl MG, Harrington MA.<br />

Developmental and tissue-specific expression of mouse pelle-like protein kinase.<br />

J Biol Chem 1996;271:17609-17612.<br />

277. Natoli G, Costanzo A, Ianni A, Templeton DJ, Woodgett JR, Balsano C, Levrero<br />

M. Activation of SAPK/JNK by TNF receptor 1 through a noncytotoxic TRAF2dependent<br />

pathway. Science 1997;275:200-203.<br />

278. Minden A, Lin A, McMahon M, Lange-Carter C, Derijard B, Davis RJ, Johnson<br />

GL, et al. Differential activation of ERK and JNK mitogen-activated protein<br />

kinases by Raf-1 and MEKK. Science 1994;266:1719-1723.<br />

279. Liu ZG, Hsu H, Goeddel DV, Karin M. Dissection of TNF receptor 1 effector<br />

functions: JNK activation is not linked to apoptosis while NF-kappaB activation<br />

prevents cell death. Cell 1996;87:565-576.<br />

280. Reinhard C, Shamoon B, Shyamala V, Williams LT. Tumor necrosis factor alphainduced<br />

activation of c-jun N-terminal kinase is mediated by TRAF2. Embo J<br />

1997;16:1080-1092.<br />

281. Wajant H, Scheurich P. Tumor necrosis factor receptor-associated factor (TRAF)<br />

2 and its role in TNF signaling. Int J Biochem Cell Biol 2001;33:19-32.<br />

282. Deveraux QL, Reed JC. IAP family proteins--suppressors of apoptosis. Genes<br />

Dev 1999;13:239-252.<br />

283. Deveraux QL, Roy N, Stennicke HR, Van Arsdale T, Zhou Q, Srinivasula SM,<br />

Alnemri ES, et al. IAPs block apoptotic events induced by caspase-8 and<br />

cytochrome c by direct inhibition of distinct caspases. Embo J 1998;17:2215-<br />

2223.<br />

284. Deveraux QL, Stennicke HR, Salvesen GS, Reed JC. Endogenous inhibitors of<br />

caspases. J Clin Immunol 1999;19:388-398.<br />

285. Nishitoh H, Saitoh M, Mochida Y, Takeda K, Nakano H, Rothe M, Miyazono K, et<br />

al. ASK1 is essential for JNK/SAPK activation by TRAF2. Mol Cell 1998;2:389-<br />

395.<br />

- 154 -


References<br />

286. Derijard B, Raingeaud J, Barrett T, Wu IH, Han J, Ulevitch RJ, Davis RJ.<br />

Independent human MAP-kinase signal transduction pathways defined by MEK<br />

and MKK isoforms. Science 1995;267:682-685.<br />

287. Kyriakis JM, Avruch J. Protein kinase cascades activated by stress and<br />

inflammatory cytokines. Bioessays 1996;18:567-577.<br />

288. Winston BW, Chan ED, Johnson GL, Riches DW. Activation of p38mapk, MKK3,<br />

and MKK4 by TNF-alpha in mouse bone marrow-derived macrophages. J<br />

Immunol 1997;159:4491-4497.<br />

289. Stokoe D, Stephens LR, Copeland T, Gaffney PR, Reese CB, Painter GF,<br />

Holmes AB, et al. Dual role of phosphatidylinositol-3,4,5-trisphosphate in the<br />

activation of protein kinase B. Science 1997;277:567-570.<br />

290. Ludwig S, Engel K, Hoffmeyer A, Sithanandam G, Neufeld B, Palm D, Gaestel M,<br />

et al. 3pK, a novel mitogen-activated protein (MAP) kinase-activated protein<br />

kinase, is targeted by three MAP kinase pathways. Mol Cell Biol 1996;16:6687-<br />

6697.<br />

291. Heinrich M, Neumeyer J, Jakob M, Hallas C, Tchikov V, Winoto-Morbach S,<br />

Wickel M, et al. Cathepsin D links TNF-induced acid sphingomyelinase to Bidmediated<br />

caspase-9 and -3 activation. Cell Death Differ 2004;11:550-563.<br />

292. Anderson ME. Glutathione and glutathione delivery compounds. Adv Pharmacol<br />

1997;38:65-78.<br />

293. Meister A. On the antioxidant effects of ascorbic acid and glutathione. Biochem<br />

Pharmacol 1992;44:1905-1915.<br />

294. Jones TW, Thor H, Orrenius S. Cellular defense mechanisms against toxic<br />

substances. Arch Toxicol Suppl 1986;9:259-271.<br />

295. Droge W, Schulze-Osthoff K, Mihm S, Galter D, Schenk H, Eck HP, Roth S, et al.<br />

Functions of glutathione and glutathione disulfide in immunology and<br />

immunopathology. Faseb J 1994;8:1131-1138.<br />

296. Uhlig S, Wendel A. The physiological consequences of glutathione variations.<br />

Life Sci 1992;51:1083-1094.<br />

297. Hentze H, Kunstle G, Volbracht C, Ertel W, Wendel A. CD95-Mediated murine<br />

hepatic apoptosis requires an intact glutathione status. Hepatology 1999;30:177-<br />

185.<br />

298. Hentze H, Gantner F, Kolb SA, Wendel A. Depletion of hepatic glutathione<br />

prevents death receptor-dependent apoptotic and necrotic liver injury in mice. Am<br />

J Pathol 2000;156:2045-2056.<br />

299. McCloskey P, Edwards RJ, Tootle R, Selden C, Roberts E, Hodgson HJ.<br />

Resistance of three immortalized human hepatocyte cell lines to acetaminophen<br />

and N-acetyl-p-benzoquinoneimine toxicity. J Hepatol 1999;31:841-851.<br />

- 155 -


References<br />

300. Abdel Rahim AG, Arthur JR, Mills CF. Effects of dietary copper, cadmium, iron,<br />

molybdenum and manganese on selenium utilization by the rat. J Nutr<br />

1986;116:403-411.<br />

301. Chuang JI, Chang TY, Liu HS. Glutathione depletion-induced apoptosis of Haras-transformed<br />

NIH3T3 cells can be prevented by melatonin. Oncogene<br />

2003;22:1349-1357.<br />

302. He YY, Huang JL, Ramirez DC, Chignell CF. Role of reduced glutathione efflux in<br />

apoptosis of immortalized human keratinocytes induced by UVA. J Biol Chem<br />

2003;278:8058-8064.<br />

303. Celli A, Que FG, Gores GJ, LaRusso NF. Glutathione depletion is associated with<br />

decreased Bcl-2 expression and increased apoptosis in cholangiocytes. Am J<br />

Physiol 1998;275:G749-G757.<br />

304. Chen Q, Galleano M, Cederbaum AI. Cytotoxicity and apoptosis produced by<br />

arachidonic acid in HepG2 cells overexpressing human cytochrome P-4502E1.<br />

Alcohol Clin Exp Res 1998;22:782-784.<br />

305. Latta M, Kunstle G, Leist M, Wendel A. Metabolic depletion of ATP by fructose<br />

inversely controls CD95- and tumor necrosis factor receptor 1-mediated hepatic<br />

apoptosis. J Exp Med 2000;191:1975-1986.<br />

306. Latta M, <strong>Kresse</strong> M, Künstle G, Lucas R, Wendel A. Models of cytokine-induced<br />

hepatic failure. Disease Progression and Carcinogenesis in the Gastrointestinal<br />

Tract (P.R. Galle, W.E. Schmidt, G. Gerken, B. Wiedenmann, eds.) Proceedings<br />

of the FALK Symposium 132 held in Freiburg, Germany, Oct. 9-10, 2002<br />

2003:86-99.<br />

307. Smith LH, Jr., Ettinger RH, Seligson D. A comparison of the metabolism of<br />

fructose and glucose in hepatic disease and diabetes mellitus. J Clin Invest<br />

1953;32:273-282.<br />

308. Miller M, Murphy JR, Craig JW, Woodward H. Studies on experimental diabetic<br />

acidosis. Comparison of effect of fructose and glucose in initial hours of<br />

treatment. J Clin Endocr 1953;13:866.<br />

309. Woods HF, Eggleston LV, Krebs HA. The cause of hepatic accumulation of<br />

fructose 1-phosphate on fructose loading. Biochem J 1970;119:501-510.<br />

310. Adelman RC, Morris HP, Weinhouse S. Fructokinase, triokinase, and aldolases in<br />

liver tumors of the rat. Cancer Res 1967;27:2408-2413.<br />

311. Bode JC, Zelder O, Rumpelt HJ, Wittkamp U. Depletion of liver adenosine<br />

phosphates and metabolic effects of intravenous infusion of fructose or sorbitol in<br />

man and in the rat. Eur J Clin Invest 1973;3:436-441.<br />

312. Pan G, Ni J, Wei YF, Yu G, Gentz R, Dixit VM. An antagonist decoy receptor and<br />

a death domain-containing receptor for TRAIL. Science 1997;277:815-818.<br />

313. Leverkus M, Neumann M, Mengling T, Rauch CT, Brocker EB, Krammer PH,<br />

Walczak H. Regulation of tumor necrosis factor-related apoptosis-inducing ligand<br />

- 156 -


References<br />

sensitivity in primary and transformed human keratinocytes. Cancer Res<br />

2000;60:553-559.<br />

314. Scaffidi C, Schmitz I, Krammer PH, Peter ME. The role of c-FLIP in modulation of<br />

CD95-induced apoptosis. J Biol Chem 1999;274:1541-1548.<br />

315. Scaffidi C, Fulda S, Srinivasan A, Friesen C, Li F, Tomaselli KJ, Debatin KM, et<br />

al. Two CD95 (APO-1/Fas) signaling pathways. Embo J 1998;17:1675-1687.<br />

316. Irmler M, Thome M, Hahne M, Schneider P, Hofmann K, Steiner V, Bodmer JL, et<br />

al. Inhibition of death receptor signals by cellular FLIP. Nature 1997;388:190-195.<br />

317. Goyal L. Cell death inhibition: keeping caspases in check. Cell 2001;104:805-<br />

808.<br />

318. Harris NL, Stein H, Coupland SE, Hummel M, Favera RD, Pasqualucci L, Chan<br />

WC. New approaches to lymphoma diagnosis. Hematology (Am Soc Hematol<br />

Educ Program) 2001:194-220.<br />

319. Cavalli F, Isaacson PG, Gascoyne RD, Zucca E. MALT Lymphomas. Hematology<br />

(Am Soc Hematol Educ Program) 2001:241-258.<br />

320. Dierlamm J, Baens M, Wlodarska I, Stefanova-Ouzounova M, Hernandez JM,<br />

Hossfeld DK, De Wolf-Peeters C, et al. The apoptosis inhibitor gene API2 and a<br />

novel 18q gene, MLT, are recurrently rearranged in the<br />

t(11;18)(q21;q21)p6ssociated with mucosa-associated lymphoid tissue<br />

lymphomas. Blood 1999;93:3601-3609.<br />

321. Gutierrez MI, Cherney B, Hussain A, Mostowski H, Tosato G, Magrath I, Bhatia<br />

K. Bax is frequently compromised in Burkitt's lymphomas with irreversible<br />

resistance to Fas-induced apoptosis. Cancer Res 1999;59:696-703.<br />

322. MacNamara B, Wang W, Chen Z, Hou M, Mazur J, Gruber A, Porwit-MacDonald<br />

A. Telomerase activity in relation to pro- and anti-apoptotic protein expression in<br />

high grade non-Hodgkin's lymphomas. Haematologica 2001;86:386-393.<br />

323. Meijerink JP, Mensink EJ, Wang K, Sedlak TW, Sloetjes AW, de Witte T,<br />

Waksman G, et al. Hematopoietic malignancies demonstrate loss-of-function<br />

mutations of BAX. Blood 1998;91:2991-2997.<br />

324. Snow AL, Chen LJ, Nepomuceno RR, Krams SM, Esquivel CO, Martinez OM.<br />

Resistance to Fas-mediated apoptosis in EBV-infected B cell lymphomas is due<br />

to defects in the proximal Fas signaling pathway. J Immunol 2001;167:5404-<br />

5411.<br />

325. Takakuwa T, Dong Z, Takayama H, Matsuzuka F, Nagata S, Aozasa K. Frequent<br />

mutations of Fas gene in thyroid lymphoma. Cancer Res 2001;61:1382-1385.<br />

326. Bertoni F, Conconi A, Luminari S, Realini C, Roggero E, Baldini L, Carobbio S, et<br />

al. Lack of CD95/FAS gene somatic mutations in extranodal, nodal and splenic<br />

marginal zone B cell lymphomas. Leukemia 2000;14:446-448.<br />

- 157 -


References<br />

327. Gronbaek K, Straten PT, Ralfkiaer E, Ahrenkiel V, Andersen MK, Hansen NE,<br />

Zeuthen J, et al. Somatic Fas mutations in non-Hodgkin's lymphoma: association<br />

with extranodal disease and autoimmunity. Blood 1998;92:3018-3024.<br />

328. Xerri L, Carbuccia N, Parc P, Birg F. Search for rearrangements and/or allelic<br />

loss of the fas/APO-1 gene in 101 human lymphomas. Am J Clin Pathol<br />

1995;104:424-430.<br />

329. Shin MS, Kim HS, Kang CS, Park WS, Kim SY, Lee SN, Lee JH, et al.<br />

Inactivating mutations of CASP10 gene in non-Hodgkin lymphomas. Blood<br />

2002;99:4094-4099.<br />

330. Dales JP, Palmerini F, Devilard E, Hassoun J, Birg F, Xerri L. Caspases:<br />

conductors of the cell death machinery in lymphoma cells. Leuk Lymphoma<br />

2001;41:247-253.<br />

331. Kuchino Y, Kitanaka C. Apoptosis in cancer. Hum Cell 1996;9:223-228.<br />

332. Arends MJ, Wyllie AH. Apoptosis: mechanisms and roles in pathology. Int Rev<br />

Exp Pathol 1991;32:223-254.<br />

333. Mesner PW, Jr., Budihardjo, II, Kaufmann SH. Chemotherapy-induced apoptosis.<br />

Adv Pharmacol 1997;41:461-499.<br />

334. Liebermann DA, Hoffman B, Steinman RA. Molecular controls of growth arrest<br />

and apoptosis: p53-dependent and independent pathways. Oncogene<br />

1995;11:199-210.<br />

335. Kato M, Sato H. Regulation of apoptosis by p53. Exp Med 1995;13:1866-1872.<br />

336. Scrable HJ, Sapienza C, Cavenee WK. Genetic and epigenetic losses of<br />

heterozygosity in cancer predisposition and progression. Adv Cancer Res<br />

1990;54:25-62.<br />

337. Sugiyama A, Kume A, Nemoto K, Lee SY, Asami Y, Nemoto F, Nishimura S, et<br />

al. Isolation and characterization of s-myc, a member of the rat myc gene family.<br />

Proc Natl Acad Sci U S A 1989;86:9144-9148.<br />

338. Fourel G, Trepo C, Bougueleret L, Henglein B, Ponzetto A, Tiollais P, Buendia<br />

MA. Frequent activation of N-myc genes by hepadnavirus insertion in woodchuck<br />

liver tumours. Nature 1990;347:294-298.<br />

339. Asai A, Miyagi Y, Sugiyama A, Nagashima Y, Kanemitsu H, Obinata M, Mishima<br />

K, et al. The s-Myc protein having the ability to induce apoptosis is selectively<br />

expressed in rat embryo chondrocytes. Oncogene 1994;9:2345-2352.<br />

340. Kitanaka C, Sugiyama A, Kanazu S, Miyagi Y, Mishima K, Asai A, Kuchino Y. s-<br />

Myc acts as a transcriptional activator and its sequence-specific DNA binding is<br />

required for induction of programmed cell death in glioma cells. Cell Death Differ<br />

1995;2:123-131.<br />

341. Kuchino Y, Kitanaka C, Asai A. [Application of myc family genes to therapy of<br />

malignant brain tumors]. Tanpakushitsu Kakusan Koso 1995;40:2713-2719.<br />

- 158 -


References<br />

342. Kaufmann SH, Earnshaw WC. Induction of apoptosis by cancer chemotherapy.<br />

Exp Cell Res 2000;256:42-49.<br />

343. Friesen C, Fulda S, Debatin KM. Deficient activation of the CD95 (APO-1/Fas)<br />

system in drug-resistant cells. Leukemia 1997;11:1833-1841.<br />

344. Friesen C, Herr I, Krammer PH, Debatin KM. Involvement of the CD95 (APO-<br />

1/FAS) receptor/ligand system in drug-induced apoptosis in leukemia cells. Nat<br />

Med 1996;2:574-577.<br />

345. Kasibhatla S, Brunner T, Genestier L, Echeverri F, Mahboubi A, Green DR. DNA<br />

damaging agents induce expression of Fas ligand and subsequent apoptosis in T<br />

lymphocytes via the activation of NF-kappa B and AP-1. Mol Cell 1998;1:543-<br />

551.<br />

346. Fulda S, Sieverts H, Friesen C, Herr I, Debatin KM. The CD95 (APO-1/Fas)<br />

system mediates drug-induced apoptosis in neuroblastoma cells. Cancer Res<br />

1997;57:3823-3829.<br />

347. Fulda S, Scaffidi C, Pietsch T, Krammer PH, Peter ME, Debatin KM. Activation of<br />

the CD95 (APO-1/Fas) pathway in drug- and gamma-irradiation-induced<br />

apoptosis of brain tumor cells. Cell Death Differ 1998;5:884-893.<br />

348. Fulda S, Los M, Friesen C, Debatin KM. Chemosensitivity of solid tumor cells in<br />

vitro is related to activation of the CD95 system. Int J Cancer 1998;76:105-114.<br />

349. Bennett MW, O'Connell J, O'Sullivan GC, Brady C, Roche D, Collins JK,<br />

Shanahan F. The Fas counterattack in vivo: apoptotic depletion of tumorinfiltrating<br />

lymphocytes associated with Fas ligand expression by human<br />

esophageal carcinoma. J Immunol 1998;160:5669-5675.<br />

350. Michael-Robinson JM, Pandeya N, Cummings MC, Walsh MD, Young JP,<br />

Leggett BA, Purdie DM, et al. Fas ligand and tumour counter-attack in colorectal<br />

cancer stratified according to microsatellite instability status. J Pathol<br />

2003;201:46-54.<br />

351. Hug H, Strand S, Grambihler A, Galle J, Hack V, Stremmel W, Krammer PH, et<br />

al. Reactive oxygen intermediates are involved in the induction of CD95 ligand<br />

mRNA expression by cytostatic drugs in hepatoma cells. J Biol Chem<br />

1997;272:28191-28193.<br />

352. Jovasevic VM, Mokyr MB. Melphalan-induced expression of IFN-beta in MOPC-<br />

315 tumor-bearing mice and its importance for the up-regulation of TNF-alpha<br />

expression. J Immunol 2001;167:4895-4901.<br />

353. Houghton JA, Harwood FG, Tillman DM. Thymineless death in colon carcinoma<br />

cells is mediated via fas signaling. Proc Natl Acad Sci U S A 1997;94:8144-8149.<br />

354. Tillman DM, Petak I, Houghton JA. A Fas-dependent component in 5fluorouracil/leucovorin-induced<br />

cytotoxicity in colon carcinoma cells. Clin Cancer<br />

Res 1999;5:425-430.<br />

- 159 -


References<br />

355. McGahon AJ, Costa Pereira AP, Daly L, Cotter TG. Chemotherapeutic druginduced<br />

apoptosis in human leukaemic cells is independent of the Fas (APO-<br />

1/CD95) receptor/ligand system. Br J Haematol 1998;101:539-547.<br />

356. Eischen CM, Kottke TJ, Martins LM, Basi GS, Tung JS, Earnshaw WC, Leibson<br />

PJ, et al. Comparison of apoptosis in wild-type and Fas-resistant cells:<br />

chemotherapy-induced apoptosis is not dependent on Fas/Fas ligand<br />

interactions. Blood 1997;90:935-943.<br />

357. Shain KH, Landowski TH, Buyuksal I, Cantor AB, Dalton WS. Clonal variability in<br />

CD95 expression is the major determinant in Fas-medicated, but not<br />

chemotherapy-medicated apoptosis in the RPMI 8226 multiple myeloma cell line.<br />

Leukemia 2000;14:830-840.<br />

358. Seglen PO. Preparation of rat liver cells. 3. Enzymatic requirements for tissue<br />

dispersion. Exp Cell Res 1973;82:391-398.<br />

359. Klaunig JE, Goldblatt PJ, Hinton DE, Lipsky MM, Chacko J, Trump BF. Mouse<br />

liver cell culture. I. Hepatocyte isolation. In Vitro 1981;17:913-925.<br />

360. Klaunig JE, Goldblatt PJ, Hinton DE, Lipsky MM, Trump BF. Mouse liver cell<br />

culture. II. Primary culture. In Vitro 1981;17:926-934.<br />

361. Leist M. Inflammatory stimuli and mediators in the mouse liver. <strong>Konstanz</strong>:<br />

University of <strong>Konstanz</strong>; 1993.<br />

362. Osypiw JC, Allen RL, Billington D. Subpopulations of rat hepatocytes separated<br />

by Percoll density-gradient centrifugation show characteristics consistent with<br />

different acinar locations. Biochem J 1994;304:617-624.<br />

363. Van Rooijen N, Sanders A. Liposome mediated depletion of macrophages:<br />

mechanism of action, preparation of liposomes and applications. J Immunol<br />

Methods 1994;174:83-93.<br />

364. Doolittle RL, Richter GW. Isolation and culture of Kupffer cells and hepatocytes<br />

from single rat livers: with observations on iron-loaded kupffer cells. Lab Invest<br />

1981;45:558-566.<br />

365. Schümann J, Wolf D, Pahl A, Brune K, Papadopoulos T, van Rooijen N, Tiegs G.<br />

Importance of kupffer cells for T-cell-dependent liver injury in mice. Am J Pathol<br />

2000;157:1671-1683.<br />

366. Tiegs G, Hentschel J, Wendel A. A T cell-dependent experimental liver injury in<br />

mice inducible by concanavalin A. J Clin Invest 1992;90:196-203.<br />

367. Bergmeyer HU. Evaluation of optimum conditions of two-substrate enzyme<br />

reactions. J Clin Chem Clin Biochem 1977;15:405-410.<br />

368. Bergmeyer HU, Scheibe P, Wahlefeld AW. Optimization of methods for aspartate<br />

aminotransferase and alanine aminotransferase. Clin Chem 1978;24:58-73.<br />

369. Bergmeyer HU. Methods of enzymatic analysis. Weinheim: Verlag Chemie<br />

Weinheim, 1983: 605.<br />

- 160 -


References<br />

370. Thornberry NA. Caspases: a decade of death research. Cell Death Differ<br />

1999;6:1023-1027.<br />

371. Jin G, Huang X, Black R, Wolfson M, Rauch C, McGregor H, Ellestad G, et al. A<br />

continuous fluorimetric assay for tumor necrosis factor-alpha converting enzyme.<br />

Anal Biochem 2002;302:269-275.<br />

372. Tietze F. Enzymic method for quantitative determination of nanogram amounts of<br />

total and oxidized glutathione: applications to mammalian blood and other<br />

tissues. Anal Biochem 1969;27:502-522.<br />

373. Lucas R, Juillard P, Decoster E, Redard M, Burger D, Donati Y, Giroud C, et al.<br />

Crucial role of tumor necrosis factor (TNF) receptor 2 and membrane- bound TNF<br />

in experimental cerebral malaria. Eur J Immunol 1997;27:1719-1725.<br />

374. Lucas R, Heirwegh K, Neirynck A, Remels L, Van Heuverswyn H, De Baetselier<br />

P. Generation and characterization of a neutralizing rat anti-rmTNF-alpha<br />

monoclonal antibody. Immunology 1990;71:218-223.<br />

375. Watanabe-Fukunaga R, Brannan CI, Copeland NG, Jenkins NA, Nagata S.<br />

Lymphoproliferation disorder in mice explained by defects in Fas antigen that<br />

mediates apoptosis. Nature 1992;356:314-317.<br />

376. Vanhaesebroeck B, Van Bladel S, Lenaerts A, Suffys P, Beyaert R, Lucas R, Van<br />

Roy F, et al. Two discrete types of tumor necrosis factor-resistant cells derived<br />

from the same cell line. Cancer Res 1991;51:2469-2477.<br />

377. Jones AL, Selby P. Tumour necrosis factor: clinical relevance. Cancer Surv<br />

1989;8:817-836.<br />

378. Lehmann V, Freudenberg MA, Galanos C. Lethal toxicity of lipopolysaccharide<br />

and tumor necrosis factor in normal and D-galactosamine-treated mice. J Exp<br />

Med 1987;165:657-663.<br />

379. Tiegs G, Wolter M, Wendel A. Tumor necrosis factor is a terminal mediator in<br />

galactosamine/endotoxin-induced hepatitis in mice. Biochem Pharmacol<br />

1989;38:627-631.<br />

380. Decker K, Keppler D. Galactosamine hepatitis: key role of the nucleotide<br />

deficiency period in the pathogenesis of cell injury and cell death. Rev Physiol<br />

Biochem Pharmacol 1974;71:77-106.<br />

381. Leist M, Auer-Barth S, Wendel A. Tumor necrosis factor production in the<br />

perfused mouse liver and its pharmacological modulation by methylxanthines. J<br />

Pharmacol Exp Ther 1996;176:968-976.<br />

382. Gantner F, Leist M, Kusters S, Vogt K, Volk HD, Tiegs G. T cell stimulus-induced<br />

crosstalk between lymphocytes and liver macrophages results in augmented<br />

cytokine release. Exp Cell Res 1996;229:137-146.<br />

383. Leist M, Gantner F, Bohlinger I, Tiegs G, Germann PG, Wendel A. Tumor<br />

necrosis factor-induced hepatocyte apoptosis precedes liver failure in<br />

experimental murine shock models. Am J Pathol 1995;146:1220-1234.<br />

- 161 -


References<br />

384. Tiegs G, Niehorster M, Wendel A. Leukocyte alterations do not account for<br />

hepatitis induced by endotoxin or TNF alpha in galactosamine-sensitized mice.<br />

Biochem Pharmacol 1990;40:1317-1322.<br />

385. Schümann J, Bluethmann H, Tiegs G. Synergism of Pseudomonas aeruginosa<br />

exotoxin A with endotoxin, superantigen, or TNF results in TNFR1- and TNFR2dependent<br />

liver toxicity in mice. Immunol Lett 2000;74:165-172.<br />

386. Gantner F, Leist M, Jilg S, Germann PG, Freudenberg MA, Tiegs G. Tumor<br />

necrosis factor-induced hepatic DNA fragmentation as an early marker of T celldependent<br />

liver injury in mice. Gastroenterology 1995;109:166-176.<br />

387. Miethke T, Wahl C, Heeg K, Echtenacher B, Krammer PH, Wagner H. T cellmediated<br />

lethal shock triggered in mice by the superantigen staphylococcal<br />

enterotoxin B: critical role of tumor necrosis factor. J Exp Med 1992;175:91-98.<br />

388. Pfeffer K, Matsuyama T, Kundig TM, Wakeham A, Kishihara K, Shahinian A,<br />

Wiegmann K, et al. Mice deficient for the 55 kd tumor necrosis factor receptor are<br />

resistant to endotoxic shock, yet succumb to L. monocytogenes infection. Cell<br />

1993;73:457-467.<br />

389. Nagaki M, Muto Y, Ohnishi H, Yasuda S, Sano K, Naito T, Maeda T, et al.<br />

Hepatic injury and lethal shock in galactosamine-sensitized mice induced by the<br />

superantigen staphylococcal enterotoxin B. Gastroenterology 1994;106:450-458.<br />

390. Schümann J, Angermuller S, Bang R, Lohoff M, Tiegs G. Acute hepatotoxicity of<br />

Pseudomonas aeruginosa exotoxin A in mice depends on T cells and TNF. J<br />

Immunol 1998;161:5745-5754.<br />

391. Koesling M, Rott O, Fleischer B. Macrophages are dispensable for superantigenmediated<br />

stimulation and anergy induction of peripheral T cells in vivo. Cell<br />

Immunol 1994;157:29-37.<br />

392. Bette M, Schafer MK, van Rooijen N, Weihe E, Fleischer B. Distribution and<br />

kinetics of superantigen-induced cytokine gene expression in mouse spleen. J<br />

Exp Med 1993;178:1531-1539.<br />

393. Gantner F, Leist M, Lohse AW, Germann PG, Tiegs G. Concanavalin A-induced<br />

T-cell-mediated hepatic injury in mice: the role of tumor necrosis factor.<br />

Hepatology 1995;21:190-198.<br />

394. Toyabe S, Seki S, Iiai T, Takeda K, Shirai K, Watanabe H, Hiraide H, et al.<br />

Requirement of IL-4 and liver NK1+ T cells for concanavalin A-induced hepatic<br />

injury in mice. J Immunol 1997;159:1537-1542.<br />

395. Knolle PA, Gerken G, Loser E, Dienes HP, Gantner F, Tiegs G, Meyer zum<br />

Buschenfelde KH, et al. Role of sinusoidal endothelial cells of the liver in<br />

concanavalin A-induced hepatic injury in mice. Hepatology 1996;24:824-829.<br />

396. Santucci L, Fiorucci S, Cammilleri F, Servillo G, Federici B, Morelli A. Galectin-1<br />

exerts immunomodulatory and protective effects on concanavalin A-induced<br />

hepatitis in mice. Hepatology 2000;31:399-406.<br />

- 162 -


References<br />

397. Fiorucci S, Santucci L, Antonelli E, Distrutti E, Del Sero G, Morelli O, Romani L,<br />

et al. NO-aspirin protects from T cell-mediated liver injury by inhibiting caspasedependent<br />

processing of Th1-like cytokines. Gastroenterology 2000;118:404-<br />

421.<br />

398. Ogasawara J, Watanabe-Fukunaga R, Adachi M, Matsuzawa A, Kasugai T,<br />

Kitamura Y, Itoh N, et al. Lethal effect of the anti-Fas antibody in mice. Nature<br />

1993;364:806-809.<br />

399. Enari M, Hug H, Nagata S. Involvement of an ICE-like protease in Fas-mediated<br />

apoptosis. Nature 1995;375:78-81.<br />

400. Rensing-Ehl A, Hess S, Ziegler-Heitbrock HW, Riethmuller G, Engelmann H.<br />

Fas/Apo-1 activates nuclear factor kappa B and induces interleukin-6 production.<br />

J Inflamm 1995;45:161-174.<br />

401. Kondo T, Suda T, Fukuyama H, Adachi M, Nagata S. Essential roles of the Fas<br />

ligand in the development of hepatitis. Nat Med 1997;3:409-413.<br />

402. Nishimura Y, Hirabayashi Y, Matsuzaki Y, Musette P, Ishii A, Nakauchi H, Inoue<br />

T, et al. In vivo analysis of Fas antigen-mediated apoptosis: effects of agonistic<br />

anti-mouse Fas mAb on thymus, spleen and liver. Int Immunol 1997;9:307-316.<br />

403. Ichikawa K, Yoshida-Kato H, Ohtsuki M, Ohsumi J, Yamaguchi J, Takahashi S,<br />

Tani Y, et al. A novel murine anti-human Fas mAb which mitigates<br />

lymphadenopathy without hepatotoxicity. Int Immunol 2000;12:555-562.<br />

404. Kakinuma C, Takagaki K, Yatomi T, Nakamura N, Nagata S, Uemura A,<br />

Shibutani Y. Acute toxicity of an anti-Fas antibody in mice. Toxicol Pathol<br />

1999;27:412-420.<br />

405. Cardier JE, Schulte T, Kammer H, Kwak J, Cardier M. Fas (CD95, APO-1)<br />

antigen expression and function in murine liver endothelial cells: implications for<br />

the regulation of apoptosis in liver endothelial cells. Faseb J 1999;13:1950-1960.<br />

406. Wanner GA, Mica L, Wanner-Schmid E, Kolb SA, Hentze H, Trentz O, Ertel W.<br />

Inhibition of caspase activity prevents CD95-mediated hepatic microvascular<br />

perfusion failure and restores Kupffer cell clearance capacity. Faseb J<br />

1999;13:1239-1248.<br />

407. Wolf D, Schumann J, Koerber K, Kiemer AK, Vollmar AM, Sass G, Papadopoulos<br />

T, et al. Low-molecular-weight hyaluronic acid induces nuclear factor-kappaB-<br />

dependent resistance against tumor necrosis factor alpha-mediated liver injury in<br />

mice. Hepatology 2001;34:535-547.<br />

408. Alexander HR, Jr., Bartlett DL, Libutti SK. Current status of isolated hepatic<br />

perfusion with or without tumor necrosis factor for the treatment of unresectable<br />

cancers confined to liver. Oncologist 2000;5:416-424.<br />

409. Choti MA, Bulkley GB. Management of hepatic metastases. Liver Transpl Surg<br />

1999;5:65-80.<br />

- 163 -


References<br />

410. Venook AP. Update on hepatic intra-arterial chemotherapy. Oncology (Huntingt)<br />

1997;11:947-957; discussion 961-942, 964, 970.<br />

411. Cancer M-AGi. Reappraisal of hepatic arterial infusion in the treatment of<br />

nonresectable liver metastases from colorectal cancer. Meta-Analysis Group in<br />

Cancer. J Natl Cancer Inst 1996;88:252-258.<br />

412. Lienard D, Eggermont AM, Koops HS, Kroon B, Towse G, Hiemstra S, Schmitz<br />

P, et al. Isolated limb perfusion with tumour necrosis factor-alpha and melphalan<br />

with or without interferon-gamma for the treatment of in-transit melanoma<br />

metastases: a multicentre randomized phase II study. Melanoma Res<br />

1999;9:491-502.<br />

413. Wu PC, McCart A, Hewitt SM, Turner E, Libutti SK, Bartlett DL, Alexander HR.<br />

Isolated organ perfusion does not result in systemic microembolization of tumor<br />

cells. Ann Surg Oncol 1999;6:658-663.<br />

414. Spriggs DR, Sherman ML, Michie H, Arthur KA, Imamura K, Wilmore D, Frei E,<br />

3rd, et al. Recombinant human tumor necrosis factor administered as a 24-hour<br />

intravenous infusion. A phase I and pharmacologic study. J Natl Cancer Inst<br />

1988;80:1039-1044.<br />

415. Rudiger HA, Clavien PA. Tumor necrosis factor alpha, but not Fas, mediates<br />

hepatocellular apoptosis in the murine ischemic liver. Gastroenterology<br />

2002;122:202-210.<br />

416. Leist M, Gantner F, Bohlinger I, Germann PG, Tiegs G, Wendel A. Murine<br />

hepatocyte apoptosis induced in vitro and in vivo by TNF-alpha requires<br />

transcriptional arrest. J Immunol 1994;153:1778-1788.<br />

417. Yamada Y, Fausto N. Deficient liver regeneration after carbon tetrachloride injury<br />

in mice lacking type 1 but not type 2 tumor necrosis factor receptor. Am J Pathol<br />

1998;152:1577-1589.<br />

418. Yamada Y, Webber EM, Kirillova I, Peschon JJ, Fausto N. Analysis of liver<br />

regeneration in mice lacking type 1 or type 2 tumor necrosis factor receptor:<br />

requirement for type 1 but not type 2 receptor. Hepatology 1998;28:959-970.<br />

419. Knight B, Yeoh GC, Husk KL, Ly T, Abraham LJ, Yu C, Rhim JA, et al. Impaired<br />

preneoplastic changes and liver tumor formation in tumor necrosis factor receptor<br />

type 1 knockout mice. J Exp Med 2000;192:1809-1818.<br />

420. Luster MI, Simeonova PP, Gallucci RM, Bruccoleri A, Blazka ME, Yucesoy B,<br />

Matheson JM. The role of tumor necrosis factor alpha in chemical-induced<br />

hepatotoxicity. Ann N Y Acad Sci 2000;919:214-220.<br />

421. Simeonova PP, Gallucci RM, Hulderman T, Wilson R, Kommineni C, Rao M,<br />

Luster MI. The role of tumor necrosis factor-alpha in liver toxicity, inflammation,<br />

and fibrosis induced by carbon tetrachloride. Toxicol Appl Pharmacol<br />

2001;177:112-120.<br />

- 164 -


References<br />

422. Enomoto N, Takei Y, Hirose M, Ikejima K, Miwa H, Kitamura T, Sato N.<br />

Thalidomide prevents alcoholic liver injury in rats through suppression of Kupffer<br />

cell sensitization and TNF-alpha production. Gastroenterology 2002;123:291-300.<br />

423. Sharma RP, Bhandari N, Riley RT, Voss KA, Meredith FI. Tolerance to fumonisin<br />

toxicity in a mouse strain lacking the P75 tumor necrosis factor receptor.<br />

Toxicology 2000;143:183-194.<br />

424. Probert L, Akassoglou K, Kassiotis G, Pasparakis M, Alexopoulou L, Kollias G.<br />

TNF-alpha transgenic and knockout models of CNS inflammation and<br />

degeneration. J Neuroimmunol 1997;72:137-141.<br />

425. Akassoglou K, Douni E, Bauer J, Lassmann H, Kollias G, Probert L. Exclusive<br />

tumor necrosis factor (TNF) signaling by the p75TNF receptor triggers<br />

inflammatory ischemia in the CNS of transgenic mice. Proc Natl Acad Sci U S A<br />

2003;100:709-714.<br />

426. Akassoglou K, Probert L, Kontogeorgos G, Kollias G. Astrocyte-specific but not<br />

neuron-specific transmembrane TNF triggers inflammation and degeneration in<br />

the central nervous system of transgenic mice. J Immunol 1997;158:438-445.<br />

427. Wajant H, Pfizenmaier K, Scheurich P. Tumor necrosis factor signaling. Cell<br />

Death Differ 2003;10:45-65.<br />

428. Alexopoulou L, Pasparakis M, Kollias G. A murine transmembrane tumor<br />

necrosis factor (TNF) transgene induces arthritis by cooperative p55/p75 TNF<br />

receptor signaling. Eur J Immunol 1997;27:2588-2592.<br />

429. Douni E, Kollias G. A critical role of the p75 tumor necrosis factor receptor<br />

(p75TNF-R) in organ inflammation independent of TNF, lymphotoxin alpha, or the<br />

p55TNF-R. J Exp Med 1998;188:1343-1352.<br />

430. Grell M, Zimmermann G, Hulser D, Pfizenmaier K, Scheurich P. TNF receptors<br />

TR60 and TR80 can mediate apoptosis via induction of distinct signal pathways.<br />

J Immunol 1994;153:1963-1972.<br />

431. Vilcek J, Feldmann M. Historical review: Cytokines as therapeutics and targets of<br />

therapeutics. Trends Pharmacol Sci 2004;25:201-209.<br />

432. MacEwan DJ. TNF receptor subtype signalling: differences and cellular<br />

consequences. Cell Signal 2002;14:477-492.<br />

433. MacEwan DJ. TNF ligands and receptors--a matter of life and death. Br J<br />

Pharmacol 2002;135:855-875.<br />

434. Garcia ST, McQuillan A, Panasci L. Correlation between the cytotoxicity of<br />

melphalan and DNA crosslinks as detected by the ethidium bromide fluorescence<br />

assay in the F1 variant of B16 melanoma cells. Biochem Pharmacol<br />

1988;37:3189-3192.<br />

435. Frankfurt OS. Detection of DNA damage in individual cells by flow cytometric<br />

analysis using anti-DNA monoclonal antibody. Exp Cell Res 1987;170:369-380.<br />

- 165 -


References<br />

436. Masta A, Gray PJ, Phillips DR. Nitrogen mustard inhibits transcription and<br />

translation in a cell free system. Nucleic Acids Res 1995;23:3508-3515.<br />

437. Grell M, Wajant H, Zimmermann G, Scheurich P. The type 1 receptor (CD120a)<br />

is the high-affinity receptor for soluble tumor necrosis factor. Proc Natl Acad Sci<br />

U S A 1998;95:570-575.<br />

438. Moss ML, Jin SL, Milla ME, Bickett DM, Burkhart W, Carter HL, Chen WJ, et al.<br />

Cloning of a disintegrin metalloproteinase that processes precursor tumournecrosis<br />

factor-alpha. Nature 1997;385:733-736.<br />

439. Tiegs G. Immunotoxicology of host-response-mediated experimental liver injury.<br />

J Hepatol 1994;21:890-903.<br />

440. Nakama T, Hirono S, Moriuchi A, Hasuike S, Nagata K, Hori T, Ido A, et al.<br />

Etoposide prevents apoptosis in mouse liver with D-<br />

galactosamine/lipopolysaccharide-induced fulminant hepatic failure resulting in<br />

reduction of lethality. Hepatology 2001;33:1441-1450.<br />

441. Mizuhara H, O'Neill E, Seki N, Ogawa T, Kusunoki C, Otsuka K, Satoh S, et al. T<br />

cell activation-associated hepatic injury: mediation by tumor necrosis factors and<br />

protection by interleukin 6. J Exp Med 1994;179:1529-1537.<br />

442. Aversa G, Punnonen J, de Vries JE. The 26-kD transmembrane form of tumor<br />

necrosis factor alpha on activated CD4+ T cell clones provides a costimulatory<br />

signal for human B cell activation. J Exp Med 1993;177:1575-1585.<br />

443. Sass G, Koerber K, Bang R, Guehring H, Tiegs G. Inducible nitric oxide synthase<br />

is critical for immune-mediated liver injury in mice. J Clin Invest 2001;107:439-<br />

447.<br />

444. Ulich TR, Guo KZ, Irwin B, Remick DG, Davatelis GN. Endotoxin-induced<br />

cytokine gene expression in vivo. II. Regulation of tumor necrosis factor and<br />

interleukin-1 alpha/beta expression and suppression. Am J Pathol<br />

1990;137:1173-1185.<br />

445. Kiemer AK, Vollmar AM, Bilzer M, Gerwig T, Gerbes AL. Atrial natriuretic peptide<br />

reduces expression of TNF-alpha mRNA during reperfusion of the rat liver upon<br />

decreased activation of NF-kappaB and AP-1. J Hepatol 2000;33:236-246.<br />

446. Kiemer AK, Vollmar AM. Effects of different natriuretic peptides on nitric oxide<br />

synthesis in macrophages. Endocrinology 1997;138:4282-4290.<br />

447. de Vries MR, Borel Rinkes IH, Swaak AJ, Hack CE, Van De Velde CJ, Wiggers<br />

T, Tollenaar RA, et al. Acute-phase response patterns in isolated hepatic<br />

perfusion with tumour necrosis factor alpha (TNF-alpha) and melphalan in<br />

patients with colorectal liver metastases. Eur J Clin Invest 1999;29:553-560.<br />

448. Gorelik L, Rubin M, Prokhorova A, Mokyr MB. Importance of TNF production for<br />

the curative effectiveness of low dose melphalan therapy for mice bearing a large<br />

MOPC-315 tumor. J Immunol 1995;154:3941-3951.<br />

- 166 -


References<br />

449. Manusama ER, Nooijen PT, Stavast J, Durante NM, Marquet RL, Eggermont AM.<br />

Synergistic antitumour effect of recombinant human tumour necrosis factor alpha<br />

with melphalan in isolated limb perfusion in the rat. Br J Surg 1996;83:551-555.<br />

450. Furrer M, Altermatt HJ, Ris HB, Althaus U, Ruegg C, Lienard D, Lejeune FJ. Lack<br />

of antitumour activity of human recombinant tumour necrosis factor-alpha, alone<br />

or in combination with melphalan in a nude mouse human melanoma xenograft<br />

system. Melanoma Res 1997;7 Suppl 2:S43-S49.<br />

451. Van Molle W, Wielockx B, Mahieu T, Takada M, Taniguchi T, Sekikawa K, Libert<br />

C. HSP70 protects against TNF-induced lethal inflammatory shock. Immunity<br />

2002;16:685-695.<br />

452. Kuppen PJ, Jonges LE, van de Velde CJ, Vahrmeijer AL, Tollenaar RA, Borel<br />

Rinkes IH, Eggermont AM. Liver and tumour tissue concentrations of TNF-alpha<br />

in cancer patients treated with TNF-alpha and melphalan by isolated liver<br />

perfusion. Br J Cancer 1997;75:1497-1500.<br />

453. Lans TE, Bartlett DL, Libutti SK, Gnant MF, Liewehr DJ, Venzon DJ, Turner EM,<br />

et al. Role of tumor necrosis factor on toxicity and cytokine production after<br />

isolated hepatic perfusion. Clin Cancer Res 2001;7:784-790.<br />

454. Brinkman BM, Telliez JB, Schievella AR, Lin LL, Goldfeld AE. Engagement of<br />

tumor necrosis factor (TNF) receptor 1 leads to ATF-2- and p38 mitogenactivated<br />

protein kinase-dependent TNF-alpha gene expression. J Biol Chem<br />

1999;274:30882-30886.<br />

455. Beg AA, Baltimore D. An essential role for NF-kappaB in preventing TNF-alphainduced<br />

cell death. Science 1996;274:782-784.<br />

456. Natoli G, Costanzo A, Guido F, Moretti F, Bernardo A, Burgio VL, Agresti C, et al.<br />

Nuclear factor kB-independent cytoprotective pathways originating at tumor<br />

necrosis factor receptor-associated factor 2. J Biol Chem 1998;273:31262-31272.<br />

457. Van Antwerp DJ, Martin SJ, Verma IM, Green DR. Inhibition of TNF-induced<br />

apoptosis by NF-kappa B. Trends Cell Biol 1998;8:107-111.<br />

458. Seitz C, Muller P, Krieg RC, Mannel DN, Hehlgans T. A novel p75TNF receptor<br />

isoform mediating NFkappa B activation. J Biol Chem 2001;276:19390-19395.<br />

459. Paul A, Wilson S, Belham CM, Robinson CJ, Scott PH, Gould GW, Plevin R.<br />

Stress-activated protein kinases: activation, regulation and function. Cell Signal<br />

1997;9:403-410.<br />

460. Donnahoo KK, Shames BD, Harken AH, Meldrum DR. Review article: the role of<br />

tumor necrosis factor in renal ischemia-reperfusion injury. J Urol 1999;162:196-<br />

203.<br />

461. Meldrum DR, Donnahoo KK. Role of TNF in mediating renal insufficiency<br />

following cardiac surgery: evidence of a postbypass cardiorenal syndrome. J<br />

Surg Res 1999;85:185-199.<br />

- 167 -


References<br />

462. Dent P, Yacoub A, Contessa J, Caron R, Amorino G, Valerie K, Hagan MP, et al.<br />

Stress and radiation-induced activation of multiple intracellular signaling<br />

pathways. Radiat Res 2003;159:283-300.<br />

463. Takeda K, Matsuzawa A, Nishitoh H, Ichijo H. Roles of MAPKKK ASK1 in stressinduced<br />

cell death. Cell Struct Funct 2003;28:23-29.<br />

464. Bennett BL, Sasaki DT, Murray BW, O'Leary EC, Sakata ST, Xu W, Leisten JC,<br />

et al. SP600125, an anthrapyrazolone inhibitor of Jun N-terminal kinase. Proc<br />

Natl Acad Sci U S A 2001;98:13681-13686.<br />

465. Gupta S, Stravitz RT, Dent P, Hylemon PB. Down-regulation of cholesterol<br />

7alpha-hydroxylase (CYP7A1) gene expression by bile acids in primary rat<br />

hepatocytes is mediated by the c-Jun N-terminal kinase pathway. J Biol Chem<br />

2001;276:15816-15822.<br />

466. Han Z, Boyle DL, Chang L, Bennett B, Karin M, Yang L, Manning AM, et al. c-Jun<br />

N-terminal kinase is required for metalloproteinase expression and joint<br />

destruction in inflammatory arthritis. J Clin Invest 2001;108:73-81.<br />

467. Onishi I, Tani T, Hashimoto T, Shimizu K, Yagi M, Yamamoto K, Yoshioka K.<br />

Activation of c-Jun N-terminal kinase during ischemia and reperfusion in mouse<br />

liver. FEBS Lett 1997;420:201-204.<br />

468. Zwacka RM, Zhang Y, Zhou W, Halldorson J, Engelhardt JF.<br />

Ischemia/reperfusion injury in the liver of BALB/c mice activates AP-1 and<br />

nuclear factor kappaB independently of IkappaB degradation. Hepatology<br />

1998;28:1022-1030.<br />

469. Hentze H, Schmitz I, Latta M, Krueger A, Krammer PH, Wendel A. Glutathione<br />

dependence of caspase-8 activation at the death-inducing signaling complex. J<br />

Biol Chem 2002;222:5588-5595.<br />

470. Awasthi S, Bajpai KK, Piper JT, Singhal SS, Ballatore A, Seifert WE, Jr., Awasthi<br />

YC, et al. Interactions of melphalan with glutathione and the role of glutathione Stransferase.<br />

Drug Metab Dispos 1996;24:371-374.<br />

471. Troyano A, Fernandez C, Sancho P, de Blas E, Aller P. Effect of glutathione<br />

depletion on antitumor drug toxicity (apoptosis and necrosis) in U-937 human<br />

promonocytic cells. The role of intracellular oxidation. J Biol Chem<br />

2001;276:47107-47115.<br />

472. Calvert P, Yao KS, Hamilton TC, O'Dwyer PJ. Clinical studies of reversal of drug<br />

resistance based on glutathione. Chem Biol Interact 1998;111-112:213-224.<br />

473. Eguchi Y, Shimizu S, Tsujimoto Y. Intracellular ATP levels determine cell death<br />

fate by apoptosis or necrosis. Cancer Res 1997;57:1835-1840.<br />

474. Leist M, Single B, Castoldi AF, Kuhnle S, Nicotera P. Intracellular adenosine<br />

triphosphate (ATP) concentration: a switch in the decision between apoptosis and<br />

necrosis. J Exp Med 1997;185:1481-1486.<br />

- 168 -


References<br />

475. Liu X, Kim CN, Yang J, Jemmerson R, Wang X. Induction of apoptotic program in<br />

cell-free extracts: requirement for dATP and cytochrome c. Cell 1996;86:147-157.<br />

476. Stridh H, Fava E, Single B, Nicotera P, Orrenius S, Leist M. Tributyltin-induced<br />

apoptosis requires glycolytic adenosine trisphosphate production. Chem Res<br />

Toxicol 1999;12:874-882.<br />

477. Leist M, Single B, Naumann H, Fava E, Simon B, Kuhnle S, Nicotera P. Nitric<br />

oxide inhibits execution of apoptosis at two distinct ATP- dependent steps<br />

upstream and downstream of mitochondrial cytochrome c release. Biochem<br />

Biophys Res Commun 1999;258:215-221.<br />

478. Leist M, Single B, Naumann H, Fava E, Simon B, Kuhnle S, Nicotera P. Inhibition<br />

of mitochondrial ATP generation by nitric oxide switches apoptosis to necrosis.<br />

Exp Cell Res 1999;249:396-403.<br />

479. Ferrari D, Stepczynska A, Los M, Wesselborg S, Schulze-Osthoff K. Differential<br />

regulation and ATP requirement for caspase-8 and caspase-3 activation during<br />

CD95- and anticancer drug-induced apoptosis. J Exp Med 1998;188:979-984.<br />

480. Volbracht C, Leist M, Nicotera P. ATP controls neuronal apoptosis triggered by<br />

microtubule breakdown or potassium deprivation. Mol Med 1999;5:477-489.<br />

- 169 -

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!