apoptosis by c-myc - helsinki

102
Helsinki University Biomedical Dissertations No.192 Apoptosis by c-Myc: coupling metabolic reprogramming and mitochondrial apoptosis Anni Viheriäranta née Nieminen Biochemistry and Developmental Biology, Institute of Biomedicine, Faculty of Medicine & Research Programs Unit, Translational Cancer Biology, University of Helsinki Finland Helsinki Biomedicum Graduate Program ACADEMIC DISSERTATION To be publicly discussed, with the permission of Faculty of Medicine, University of Helsinki, in Biomedicum Helsinki Lecture hall 2, on March 21 st at 12 o´clock noon Helsinki 2014

Upload: others

Post on 12-Jul-2022

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Apoptosis by c-Myc - Helsinki

Helsinki University Biomedical Dissertations No.192

Apoptosis by c-Myc:

coupling metabolic reprogramming

and mitochondrial apoptosis

Anni Viheriäranta née Nieminen

Biochemistry and Developmental Biology,

Institute of Biomedicine, Faculty of Medicine &

Research Programs Unit, Translational Cancer Biology,

University of Helsinki

Finland

Helsinki Biomedicum Graduate Program

ACADEMIC DISSERTATION

To be publicly discussed, with the permission of Faculty of Medicine,

University of Helsinki, in Biomedicum Helsinki Lecture hall 2, on March 21st at

12 o´clock noon

Helsinki 2014

Page 2: Apoptosis by c-Myc - Helsinki

Supervisor: Juha Klefström, PhD, Research Director Translational Cancer Biology Research Program & Institute of Biomedicine Faculty of Medicine, University of Helsinki Helsinki, Finland Reviewers appointed by the Faculty: John Eriksson, PhD, Professor Ville Hietakangas,

Institute of Biotechnology University of Helsinki

Helsinki, Finland Thesis follow-up committee: Phillippe Juin, PhD, Professor John Eriksson, PhD, Professor Institut de Recherche Thérapeutique Université de Nantes France Opponent appointed by the Faculty: Martin Eilers, Biozentrum der Universität Würzburg Theodor Boveri Institut Würzburg, Germany

ISBN 978-952-10-9784-3 (nid.) ISBN 978-952-10-9785-0 (PDF) ISSN: 1457-8433 http://ethesis.helsinki.fi Yliopistopaino Helsinki 2014

Page 3: Apoptosis by c-Myc - Helsinki

To my family

Page 4: Apoptosis by c-Myc - Helsinki

Table of Contents ORIGINAL PUBLICATIONS .......................................................................................... 6

ABBREVIATIONS ............................................................................................................. 7

ABSTRACT ...................................................................................................................... 10

INTRODUCTION ............................................................................................................ 12

REVIEW OF THE LITERATURE ................................................................................ 13 1. Molecular mechanism of cancer ....................................................................................... 13

1.1 Hallmarks of cancer ........................................................................................................... 13 1.2 Excessive proliferation signals from oncogenes is guarded by tumor suppressors .......... 15

1.2.1 Mechanism of oncogene function ............................................................................................ 15 1.2.2 Tumor suppressors ................................................................................................................... 17

1.2.2.1 Master regulator p53 ....................................................................................................... 18 1.2.2.2 Tumor suppression by Lkb1 ............................................................................................. 21 1.2.2.3 AMPK as tumor suppressor ............................................................................................. 23

1.2.3 Cancer metabolism ................................................................................................................... 26 1.3 Breast cancer ..................................................................................................................... 29

2. Apoptosis ............................................................................................................................. 31 2.1 Apoptosis as a form of cell death ...................................................................................... 31 2.2 Apoptotic signaling molecules .......................................................................................... 32

2.2.1 Caspases as executioners ......................................................................................................... 33 2.2.2 Bcl-2 family ............................................................................................................................. 36

2.2.2.1 BH domains in the classification and function of Bcl-2 family members ........................ 36 2.2.2.4 BH3:groove interactions and conformations ................................................................... 38

2.3.1 Mitochondrial apoptosis pathway ............................................................................................ 40 2.3.1.1 MOMP and beyond .......................................................................................................... 40 2.3.1.2 Models for Bak/Bax activation by the Bcl-2 family proteins ........................................... 41 2.3.1.3 p53-dependent apoptotic pathway ................................................................................... 43

2.3.2 Death receptor apoptosis pathway ........................................................................................... 45 2.3.2.1 Death receptors and the DICS complex ........................................................................... 45

3. Oncogene Myc pathways regulating cell suicide and cancer ........................................ 48 3.1 Molecular functions of Myc .............................................................................................. 48 3.2 Oncogenic potential of Myc .............................................................................................. 50

3.2.1 Proliferation and differentiation by Myc ................................................................................. 50 3.2.2 Transformation mechanisms by Myc ....................................................................................... 51

3.3 Myc apoptosis .................................................................................................................... 52 3.4 Myc-induced metabolic reprogramming ........................................................................... 55

AIMS OF THE STUDY ................................................................................................... 58

Page 5: Apoptosis by c-Myc - Helsinki

MATERIALS AND METHODS ..................................................................................... 59

RESULTS AND DISCUSSION ....................................................................................... 72 4. The mitochondrial apoptosis pathway is engaged by Myc in a Bak-dependent manner

independently of cell context and death stimuli ................................................................... 72 5. The mitochondrial amplification loop as a mechanism for lethal caspase activation in

death receptor signaling ....................................................................................................... 75 6. Myc-induced Bak N-terminal exposure is a conformational change preceding Bax

activation and MOMP .......................................................................................................... 76 7. BH3 mimetic reactivates mitochondrial apoptosis in Myc mammary tumors ..................... 78 8. Cytoplasmic tumor suppressor p53 is activated by Myc to prime mitochondrial

apoptosis ............................................................................................................................... 79 9. The energy sensor AMPK is behind the Myc-mediated apoptotic sensitivity (III) ............. 81 10. Bak activation coupling oncogene-mediated metabolic stress and apoptosis ................... 82

CONCLUSIONS ............................................................................................................... 85

ACKNOWLEDGEMENTS ............................................................................................. 87

REFERENCES ................................................................................................................. 89

Page 6: Apoptosis by c-Myc - Helsinki

ORIGINAL PUBLICATIONS

The thesis is based of the following original publications, which have been assigned the following roman numerals: I

II

III Nieminen A.I.*, Eskelinen V.M.*, Haikala H.M., Tervonen T.A, Yan Y., Partanen, J.I., Klefström J. (2013). Myc-induced AMPK phospho-p53 pathway activates Bak to sensitize mitochondrial apoptosis. 110, 1839-48. IV Eskelinen VM., Haikala HM, Marques E,. Saarikoski S., Laajala D., Aittokallio T., Nieminen AI**, Klefstrom. J**, BH3 mimetic ABT-737 antagonizes Myc by aggravating apoptosis in luminal mammary tumors. Submitted.

The articles are reproduced with the kind permission of copyright holders.

∗=equal contributions

**=corresponding authors, Nieminen AI also performed experiments

Page 7: Apoptosis by c-Myc - Helsinki

ABBREVIATIONS 14-3-3 stratifin 3-PG 3-phosphoglycerate; 4-OHT 4-hydroxytamoxifen 5-FU 5-fluorouracil α-KG alpha-ketoglutarate A1/Bfl-1 Bcl-2 related protein A1 ACC acetyl-CoA Carboxylase AICAR 5-aminoimidazole-4-

carboxamide ribonucleoside AIF apoptosis-inducing factor AJ adherens junction AKT v-Akt murine thymoma viral

oncogene homolog 1 AMPK 5´adenosine monophosphate-

activated protein kinase ANT adenin nucleotide translocator APAF-1 apoptosis-activating factor-1 ARF alternative reading frame,

p14ARF ASCT2 neutral amino acid transporter ATM ataxia telangiectasia mutated ATR ataxia telangiectasia and Rad3

related protein AURKA aurora kinase A Bal1 brain-specific angiogenesis

inhibitor 1 Bak Bcl-2 homologous

antagonis/Killer Bax Bcl-2 associated X protein Bcl-2 B Cell leukemia 2 Bcl-xL Bcl2-like 1, L isoform Bcl-w Bcl-2-like 2 Bfk C1orf178, chromosome 1

open reading frame 178 BH Bcl-2 homology bHLHZ basic helix-loop-helix-zipper Bik Bcl-2 Interacting killer Bid BH3 Interacting domain death

agonist BL Burkitt´s lymphoma Bim Bcl-2 interacting mediator of

cell death Bmf Bcl-2 modifying factor Bok Bcl-2 - related ovarian killer Boo/Diva Bcl2-like 10, apoptosis facilitator BRCA1 breast cancer 1, early onset CAK cyclin-activating kinase CAM cell adhesion molecule

CARD caspase recruitment domain Casp caspase, cysteine-dependent

aspartate-specific proteases CDK cyclin dependent kinase CED-3 cell death abnormality 3 CED-4 cell death protein 4 CED-9 cell death protein 9 Chk checkpoint kinaseCK casein kinase CML chronic myelogenous leukemia c-Myc Human homologue of vMyc CSN-K COP9 signalosome associated

kinase complex C-terminal carboxy-terminal Cyt c cytochrome c DCIS ductal carcinoma in situ DD death domain DED death effector domain DISC death-induced signaling

complex including DMEM dulbecco´s modified Eagles´

medium DNA deoxyribonucleic acid DNAPK DNA protein kinase DR-4 TRAIL-R1 DR-5 TRAIL-R2 EBV Epstein-Barr virus ECM extracellular matrix EDTA ethylenediamininetetraacetic

acid EGFP green fluorescent protein EGFR EGF-receptor EMT epithelial-to-mesenchymal

transition EndoG Endonuclease G ER estrogen receptor ERBB2 v-erb-b2 avian erythroblastic

leukemia viral oncogene homolog 2, HER2/Neu

ERK extracellular signal-regulated kinase

FADH2 flavin afenine dinucleotide Fas/CD95 TNF receptor superfamily,

member 6 FADD Fas-associated protein death

domain FCS fetal calf serum

Page 8: Apoptosis by c-Myc - Helsinki

FDG-PET glucose 2-(18F)-fluoro-s-deoxy-D-glucose by positron emission tomography

FLIP flice inhibitory protein G0 Gap 0 in cell cycle G1 Gap 1 in cell cycle G6P glucose-6-phosphate

GLS glutaminase GLUT4 Glucose transporter type 4 GSK3β glycogen synthase kinase 3β HAUSP herpervirus-associated

ubiquitin-specific protein, USP7

HER2/Neu human epidermal growth factor receptor 2

HIF-1 hypoxia-inducible factor-1 HIPK2 homeiodomain-interacting HK hexokinase 2 HNPCC hereditary nonpolyposis

colorectal cancer IAP inhibitor of apoptosis ICE interleukin-convertin enzyme Ig immunoglobulin IGF-1 insulin growth factor 1 IkB nuclear factor of kappa light

polypeptide gene enhancer in B-cells inhibitor

IKK inhibitor of nuclear factor kappa-B kinase

JNK c-Jun NH2-terminal kinase kDa kilodalton LDHA lactate dehydrogenase A LKB1 liver kinase B1, PAR-4 LOH loss of heterozygosity MAC mitochondrial apoptosis-

induced channel MAPK mitogen-activated protein

kinase MAX myc-associated factor X MB myc box Mcl-1 myeloid Cell leukemia 1 MDM2 mouse double minute 2,

Hdmx human homolog MIZ-1 Myc-interacting zinc finger

protein 1 MEF mouse embryonic fibroblast MOMP mitochondrial outer

membrane permeabilization mtDNA mitochondrial DNA MYC vMyc avian

myelocytomatosis viral oncogene homolog

MycERtm Myc construct, activated conditionally by 4-hydroxytamoxifen

NADH nicotinamide adenine dinucleotide

NADPH nicotinamide adenine dinucleotide phosphate

NF-κB nuclear factor kappa-light-chain-enhancer of activated B cells

N-terminal amino-terminal OMM outer mitochondrial

membrane OXPHOS oxidative phosphorylation p16INK4A cyclin-dependent kinase

inhibitor 2A p53AIP1 p53-regulated apoptosis-

inducing protein 1 PAK Serine/threonine-protein

kinase PARP Poly(ADP)ribose polymerase PBS phosphate buffered saline PET positron emission tomography PFK2 phosphofructokinase PGC-1β PPARGC1B, peroxisome

proliferator-activated receptor gamma coactivator 1-beta

PI3K phosphoinositol 3kinase PJS Peutz-Jeghers syndrome PKR protein kinase R PKM2 pyruvate kinase M2 PMSF phenylmethylsulphonyl-

fluoride PPP pentose phosphate pathway PR progesterone receptor pRb retinoblastoma protein PS phosphatidylserine PT permeability transition PTP permeability transition pore Puma p53 upregulated modulator of

apoptosis R5P ribose-5 phosphate RIP receptor-interacting protein RNAi RNA interference ROCK-1 Rho-associated, coiled-coil

containing protein kinase 1 ROS reactive oxygen species SDS-PAGE Sodium dodecyl sulfate

polyacrylamide gel electrophoresis

shRNA Short hairpin interfering RNA

Page 9: Apoptosis by c-Myc - Helsinki

SLC7A5 large neutral amino acids transporter small subunit 1

Smac/ DIABLO Direct IAP binding protein SN2 SNAT5, Solute neutral amino

acid transporter 5 SP-1 specificity-protein 1 tBid truncated Bid TAF1 facilitates chromatin

transcription complex 1 TAK1 MAP3K7, Mitogen-activated

protein kinase kinase kinase 7 TCA tricarboxylic acid TGF-α transforming growth factor-α TGF-β transforming growth factor-β TM hydrophobic domain at

COOH-terminal TNF tumor necrosis factor TNFR tumor necrosis factor receptor TNM tumor-node-metastasis

TRAIL TNF-related apoptosis inducing ligand

TRADD TNFRSF1A-associated via death domain

TRAIL TNF-related apoptosis inducing ligand

TRAIL-R TRAIL Receptor TRAP tsartrate-resistant acid

phosphatase TRIM22 tripartite motif-containing 22 TSP-1 thrombospondin-1 VDAC voltage dependent anion

channel VEGF vascular endothelial growth factor v-Myc viral homolog of MYC wt wild type XIAP X-linked IAP

Page 10: Apoptosis by c-Myc - Helsinki

ABSTRACT

Escaping apoptosis is an important, acquired capability of tumor cells and is considered a

hallmark of cancer. Furthermore, tumor suppressors and apoptotic proteins function as

natural barriers for excessive proliferation induced by oncogenes and can be often found

as mutated or lost in cancer. Understanding how tumor cell specific apoptotic pathways

are working, may guide development of future therapeutic strategies inducing apoptosis

specifically in tumor cells. One of the highly overexpressed oncogene in cancers is Myc

that is a global transcription factor regulating variety of cellular functions such as

proliferation and apoptosis.

One of the major aims in this study was to elucidate mechanisms for Myc-induced

sensitization of cells to cell death. Herein, the results presented here provide evidence that

Myc-induced apoptosis is strictly mitochondria-dependent. Furthermore, our findings

expose new evidence for Bcl-2 family member mediated control of Myc-induced

apoptosis showing that Myc expression results in activation of Bak. Interestingly, our data

suggest a novel Bak activation step, where Myc induces Bak to undergo conformational

change and expose its N-termini, however, without further homo-oligomerization or

apoptosis. Hence, we open new ideas for the current Bak/Bax activation models suggested

earlier. In addition, our data indicate that Myc expression results in phosphorylation and

accumulations of mitochondrial p53 that is priming the mitochondria for death stimuli.

Herein, these results indicate that cytoplasmic p53 has a crucial role in Myc-induced stress

responses. The data presented here also suggest that BH3 mimetic drugs can sensitize

tumor cells to Myc-apoptosis in vivo, which also highlights the role of Bcl-xL as a key

controller of Myc-dependent apoptosis in epithelial tumors. Thereby, our results point new

therapeutic possibilities that could be used in combination to treat breast cancer.

Finally, in this study we found an interesting connection between the Myc-induced

metabolic transformation and the apoptotic signaling of Myc. We suggest that Myc

expression induces activation of metabolic pathways specific for cancer cell such as

activating the glutaminolytic pathway and promoting a catabolic mode of metabolism

through increasing fatty acid oxidation. Furthermore, our results indicate that Myc-

induced decline in the ATP contents of the cell is leading to an activation of a cellular

Page 11: Apoptosis by c-Myc - Helsinki

energy sensor AMPK. Consequently, AMPK is able to induce Bak activation and

apoptotic sensitivity via phosphorylation and stabilization of p53. Importantly, our results

bring new insight into oncogenic stress and apoptotic sensitivity by Myc highlighting the

role of metabolism. All together, it will be future challenge to understand how that

knowledge of coupled metabolic stress and apoptotic sensitivity can be used for

therapeutic interventions in cancer.

Page 12: Apoptosis by c-Myc - Helsinki

INTRODUCTION

Apoptosis is one of the key mechanisms regulating tissue homeostasis and is an important

hallmark of cancer. Moreover, apoptosis is a failsafe mechanism for damaged and

uncontrolled cell behavior and when inhibited a massive proliferation of cells can begin.

Therefore, to find efficient drugs it is intensively studied how cancer cells have acquired

the capability to inhibit apoptosis. The mitochondrial apoptosis pathway is a Bcl-2 family

regulated route that amplifies death signals and activates effector caspases that finally

destroy the cell. Increasing evidence from tumor analysis, suggest that especially the Bcl-

2 family and the proteins regulating it are crucial targets for mutations and expression

changes in cancers, leading to apoptotic abrogation.

Oncogene activation and deficiency in the tumor suppressor function are the major drivers

of tumorigenic transformation. The Myc oncogene contributes to the genesis of many

human cancers and as a transcription factor it regulated almost all cell phenotypes;

proliferation, growth, metabolic reprogramming, senescence and apoptosis. Most

importantly, the apoptosis seems to be especially abrogated in Myc-expressing tumors

highlighting the importance of understanding the molecular mechanism of Myc mediated

apoptosis.

Metabolic activity is a crucial determinant of a cell’s decision to proliferate or die. The

biosynthetic and bioenergetic needs are excessively increasing in proliferating oncogene

expressing cells. This requires a metabolic transformation capacity of the cell to bypass

such energetic stress. Furthermore, it is not fully understood how metabolic pathways such

as glycolysis communicate to cell cycle and apoptotic effectors. However, it is clear that a

complex network of signaling molecules is required to integrate metabolic inputs.

Elucidating these pathways is of great importance, as metabolic aberrations and their

downstream effects are known to contribute to the formation of cancer and more

importantly, to the treatment of cancer.

Page 13: Apoptosis by c-Myc - Helsinki

REVIEW OF THE LITERATURE

Tissues and organelles are well organized and arise by dividing from pre-existing stem

cells. Cells are differentiated to their final function and stay quiescent if meant to do so.

However, transformed cells have obtains the capability of excessive proliferation and does

not anymore respond to the programmed cell death signal leading to suicide. Most likely

such cells will form a benign or malignant tumor in future. Furthermore, malignant tumor

cells can migrate through blood and lymph vessels to distant sites in the body and invades

other tissues, form a metastasis, a secondary tumor (Weinberg, 2007).

Hanahan and Weinberg defined the “hallmarks of cancer” as a certain set of physiological

alterations that tumor cells acquire during multistep tumor progression, and which are

common for all malignant cells (Hanahan and Weinberg, 2000). Firstly, “sustained

proliferative signaling” is achieved by self-sufficiency in growth signals that enable a

tumor cell to grow in the absence of stimulatory proliferation signals (Figure 1). There are

several ways to do this such as transmembrane receptor overexpression (EGFR, HER2),

ligand-independent signaling through truncated receptor domains, integrin changes in

receptors and changes in downstream signaling molecules (H-Ras) (Hunter, 1997;

Medema et al., 1993; Slamon et al., 1987). In tumor cells, the proliferation is also

maintained by the capability to “evade growth suppressors” and being insensitive to

antigrowth signal. Often, the cell cycle clock is rescheduled and the cell transits through

the G1 phase instead of entering G0 via changes in CDK:cyclin complexes-mediated

phosphorylation of the retinoblastoma protein (pRb) (Weinberg, 1995). Also, excessive

proliferation signal by oncogenes (c-myc and erbA) induce de-differentiation and thereby

promote growth. One way of releasing proliferation is to lose contact inhibition for

example via Merlin or loss of epithelial polarity proteins (Foley and Eisenman, 1999;

Partanen et al., 2009). One feature of a tumor cell is to “evade apoptosis”, a form of

programmed cell death. Apoptotic machinery can be disrupted in multiple ways in cancer

Page 14: Apoptosis by c-Myc - Helsinki

either by inhibition of pro-apoptotic proteins by mutations in tumor suppressors or

overexpression of anti-apoptotic proteins. For instance, the tumor suppressor, p53 is

inactive in approximately 50 % of cancers either by mutations or by loss (Vogelstein et al.,

2000). In addition, constitutive survival signaling by the AKT pathway in cancer leads to

inhibition of apoptosis, and loss of tumor suppressor PTEN is one of the most commonly

mutated genes affecting apoptosis (Junttila and Evan, 2009). Also, other forms of

programmed cell death, like autophagy and necrosis can be affected (Galluzzi and

Kroemer, 2008; Hanahan and Weinberg, 2011). To be able to proliferate repeatedly, a

tumor cell has to evade senescence, in which cells enter quiescence with long-term loss of

proliferative capacity. This hallmark capability is considered as “limitless replicative

potential” (Hanahan and Weinberg, 2000). The overpassing of telomere-shortening-

induced crisis is also termed immortalization and is achieved through upregulating the

expression of the telomerase enzyme (Counter et al., 1998; Wright et al., 1989).

Figure 1. Hallmarks of cancer. Capabilities that enable tumor growth and metastasis. Modified from (Hanahan and Weinberg, 2011)

In addition, growing tumor needs oxygen and nutrient supply by the vasculature.

Therefore, one hallmark of cancer is “sustained angiogeesis”, a process where new blood

Page 15: Apoptosis by c-Myc - Helsinki

vessels grow. This is achieved by molecular changes resulting in “angiogenic switch”

(Veikkola and Alitalo, 1999). Also pericytes, endothelial supporting cells, and bone

marrow-derived cells contribute to tumor angiogenesis (Hanahan and Weinberg, 2011).

Furthermore, cancer cells need to obtain a capability to “reprogram metabolism” towards

aerobic glycolysis and catabolism. Importantly, the reason why cancer kills in 90% of

cases, is that the cancer has spread; it has migrated, invaded adjacent tissue and made new

colonies, metastases (Sporn, 1996). This “tissue invasion and metastasis” is a multistep

process and is mediated by physical coupling of cells to their microenvironment by

adherence interaction via CAMs and integrins and activation of matrix-degrading

extracellular proteases (Christofori and Semb, 1999; Werb, 1997). In addition, “tumor

promoting inflammation” and “avoiding immune destruction” are key phenotypes in

tumorigenic cell. How cancers acquire these hallmarks vary significantly, both

mechanistically and chronologically, however, “genomic instability” is one feature that

enables multistep tumorigenesis. For instance, the DNA damage is accumulating in tumors

by the loss of gatekeeper protein p53 tumor suppressor resulting in DNA repair system

(Christofori and Semb, 1999; Hanahan and Weinberg, 2011).

1.2.1 Mechanism of oncogene function

Cell growth is regulated in cells by multiple ways keeping cells in balance between

proliferation and cell death and tightly dedicated to their own differentiated function.

However, excessively activated proto-oncogenes can result in uncontrolled proliferation

and transformation of cells. These oncogenes are suggested to be activated by genetic

changes such as gene amplification, activating point mutation, chromosomal translocation

and insertional mutagenesis. These modifications affect either protein expression or

structure and can lead to constitutive activation of the oncogene. More often oncogenes

are involved in signal transductions pathways and execution of mitogenic signal.

Common oncogenes include growth factors (c-Sic), Receptor tyrosine kinases (EGFR,

VEGFR, HER), Cytoplasmic tyrosine kinases (Src), cytoplasmic serine/threorine kinases

(Raf), regulatory GTPases (Ras) and transcription factors (Myc) (Weinberg, 2007).

Page 16: Apoptosis by c-Myc - Helsinki

For instance, amplification of Her2 was reported in 30% of breast cancers. This was

shown to be consequence of the amplification of an entire chromosomal segment; an

amplicon in chromosome 17q (Tal et al., 1987). Another highly amplified oncogene is c-

myc, encoding the transcription factor and originally found as homologue to the avian

myelocytomatosis (leukemia and sarcoma) retroviral oncogene (v-myc) (Alitalo et al.,

1983). c-Myc (herein Myc) belongs to the family of myc genes (c-myc, n-myc, l-myc, s-

myc and b-myc), and its expression is deregulated in approximately one-third of human

cancers, especially in adenocarcinomas (Erisman et al., 1985). In Burkitt´s lymphomas

occurs chromosomal translocation of c-myc, where myc on chromosome 8 is placed under

control of the transcription-controlling enhancer sequences of an immunoglobulin (Ig)

gene on chromosomes 2, 14 and 22 (Dalla-Favera, 1981). The most common translocation

seems to be t(8;14) (Popescu and Zimonjic, 2002). Specifically, translocation occurs in

75% of cases in the heavy chain of Ig (IgH), but also in the k- and l-chains. In addition, n-

myc is amplified in 30% of childhood neuroblastomas. In the case of chronic myelogenous

leukemia (CML), translocation is caused by fusion of two distinct reading frames resulting

in a hybrid protein. In some cases, point mutations in the coding sequence of myc have

been found in translocated alleles of BL, however, most of the myc overexpression

mutations are absent (Bhatia et al., 1993). Less studied L-Myc has been observed in small

cell lung carcinomas (Dang, 2012). However, most often Myc expression is activated

through alterations in signaling pathways that induce or repress Myc transcription.

Another similar kind of oncogenic breakpoint cluster region of Brc-Abl gene fusion

(Philadelphia chromosome) occurs in chronic myelolytic leukemia, inducing a novel

fusion protein (de Klein et al., 1982).

In addition, tyrosine kinase signaling pathways can be constitutively active (self-

sufficiency in growth signals) because of mutation or overexpression. Indeed, the Ras

oncogene was shown to be reading frame point-mutated in 20% of human tumors in a

variety of tissues. It belongs to the family of Ras oncoproteins (h-ras, k-ras and n-ras),

which are small Guanosine-5´-triphosphate hydrolase (GTPases), having enzymatic

activity required for efficient signal transduction by growth factors (Malumbres and

Barbacid, 2003). The Ras pathway induces Raf and MAP kinase signaling leading to ERK

and PI3K activation. Later it was understood that Ras requires additional oncogene

Page 17: Apoptosis by c-Myc - Helsinki

cooperation, for example, with Myc or E1A (Land et al., 1983; Newbold and Overell,

1983). Moreover, KRAS mutations are found from colon, lung and pancreatic carcinomas,

whereas HRAS from breast cancer.

1.2.2 Tumor suppressors

To avoid uncontrolled cell proliferation, vertebrates have a well-designed growth control

system and various tumor suppressive mechanisms. To discriminate accurately between

normal and neoplastic growth, sensors called tumor suppressors function in diverse

cellular activities, including cell cycle checkpoint, DNA damage repair and, for example,

in protein degradation. Mutations in tumor suppressor genes behave recessively and a

single copy of the gene is enough for protein function (Weinberg, 2007). Therefore,

neoplastic growth requires two successive somatic mutations that occur sporadically. First,

the “classic” tumor suppressor gene, retinoblastoma (Rb), which is a transcriptional

repressor of cell cycle genes in the G1 phase, was found in retinoblastoma, childhood eye

cancer (Knudson, 1971; Sherr, 2004). In this cancer, the wild-type allele of the Rb gene

can be eliminated by mitotic recombination and results in a loss of heterozygosity (LOH),

where the normal copy of the gene is lost as well (Weinberg, 2007). Another important

epigenetic mechanism for inactivating tumor suppressor genes is promoter methylation,

where CpG sequences are methylated. Examples of methylated genes in tumors are the

tumor suppressor p16INK4A that was found in smokers and the BRCA1 gene, which is lost

in familial breast and ovarian cancers. From the same locus as p16INK4A, an another tumor

suppressor gene is encoded, the alternative reading frame, ARF protein that protects p53

from MDM2-mediated degradation (Sherr, 2001). Loss of ARF co-operates with

oncogenes in tumorigenesis in mice, since the apoptotic pathway is abrogated. Another

group of tumor suppressors are originally genes known to affect development, such as

Smad4 and PTEN, and misregulation of these might induce de-differentiation and enhance

tumorigenicity (Sherr, 2004).

Importantly, the DNA damage response and genomic instability are major causes for

mutagenesis in cancer. Furthermore, DNA mismatch repair genes contribute to cancer

development and are found mutated in HNPCC colorectal cancer with microsatellite

instability. Importantly, disruption of the mismatch repair system has been suggested to

Page 18: Apoptosis by c-Myc - Helsinki

lead a “mutated” phenotype in which resulting genetic instability ultimately targets other

oncogenes and tumor suppressors (Loeb et al., 2003). In addition, several DNA damage

activated kinases, ATM (ataxia telangiectasia), CHK2, and DNA-PK, are found mutated

in cancers. Furthermore, mutations in the BRCA genes are the main cause of familial

breast and ovarian cancers (Sherr, 2004) .

Transcription factor and tumor suppressor protein TP53 (tumor protein, p53) is a critical

coordinator of wide range of stress responses. Originally, p53 was discovered in 1979

from studies on SV40 T antigen by investigators from six groups (Kress et al., 1979; Lane

and Crawford, 1979). Interestingly, p53 was first described as an oncogene, not tumor

suppressor. Recent evidence again suggest that mutated p53 drives tumorigenesis having

gain-of-function properties (Brosh and Rotter, 2009; Lane and Benchimol, 1990; Muller

and Vousden, 2013). Later on, Lane et al. described p53 as the “guardian of the genome”

and Levine et al. as the “cellular gatekeeper” demonstrating the important role of p53 in

antiproliferative responses (Lane, 1992; Levine, 1997). Furthermore, the p53 is shown to

be mutated in a variety of cancers and is currently considered a most mutated gene in

cancer since one half of human cancers express a mutant p53 (Olivier et al., 2010). In

normal unstressed cells, p53 activity is maintained at low levels through a combination of

p53 degradation in ubiquitin-proteosome, principally mediated by Mdm2 (Hdm2, human

homologue) (Wade et al., 2006). In addition, Mdm2 is transcribed by p53 itself via a

negative feedback loop and it recognizes p53 as a target that should be ubiquitylated

shortly after its synthesis and finally destroyed. This also partly explains why mutated

p53 is accumulating since p53 lose its transcription-activating powers and Mdm2 is not

transcribed anymore.

Oncogenic signaling and cellular stress (hypoxia, nutrient deprivation) or DNA damage

signals increase p53 activity, which triggers apoptosis, senescence and repair programs

(Lowe et al., 2004). In addition, variety of covalent modifications of p53 occurs when p53

is stabilized and activated, many affecting the C-terminal domain (Figure 2). These post-

translational modifications include acetylation, glycosylation, methylation,

phosphorylation, ribosylation, sumoylation and ubiquitination (Weinberg, 2007; Xu,

2003). These modifications affect the ability of p53 to interact physically with other

Page 19: Apoptosis by c-Myc - Helsinki

transcription factors that modulate its transcriptional capacity and finally enable the

transcription of target genes. Furthermore, p53 is phosphorylated at multiple sites at N-

and C-terminus by a number of kinases. For instance, phosphorylation of p53 amino acid

residues at the N-terminal transactivation domain by kinases blocks Mdm2 binding and

stabilizes p53 (Figure 2). Additionally, domains near the N-terminus contribute to the

apoptotic function of p53. Whereas, the center of p53 acts as a DNA binding domain and

are phosphorylated. Moreover, the C-terminus possesses an oligomerization domain and

when phosphorylated allows p53 to form tetramers (Xu, 2003). In addition, near the C-

terminus exist also nuclear localization signals.

Figure 2. Post-translational modifications of p53. TAD, transactivation domain; SH3, Src homology 3-like domain; NLS, nuclear localization signal; DNA binding domain; TET, tetramerization domain; NES, export signal; REG, carboxy-terminal regulatory domain. Kinases regulating major phosphorylation sites of human p53 are shown and their effect on p53 functions (pink). DNA “hot-spot” mutation area in DNA binding domain (red) and areas for other post-translational modification (blue) are shown. Modified from (Bode and Dong, 2004).

However, several mechanisms have been suggested to induce the loss of p53 function in

cancer. Indeed, p53 is deleted for example in colorectal cancer or highly mutated in

inherited cancers, such as in Li-Fraumeni syndrome, and in a variety of sporadic cancers

Page 20: Apoptosis by c-Myc - Helsinki

(Srivastava et al., 1990). Furthermore, some cancer-inducing DNA viruses, such as SV40

or HPV, encode proteins like E6 that target p53 for degradation and, thereby inactivate

p53. Alternatively, some upstream pathways of p53 are mutated in cancer (Muller and

Vousden, 2013). Normally p53 exists as a homotetramer of four identical polypeptide

subunits. Importantly, more than 75% of the mutations result in the expression of a p53

protein that has lost wild-type function and may exert dominant-negative regulation over

any remaining wild-type p53 thereby inhibit its function. Interestingly, mutant p53 also

acquires oncogenic functions, a so-called gain-of-function mutant, that is entirely

independent of wild-type p53 for instance metabolic capacities (Muller and Vousden,

2013). The primary alteration in p53 is a single amino acid substitution in the 393-amino

acid protein. Most of the mutations cluster within the central DNA-binding domain and a

number of hotspots (Figure 2) have been identified (Muller and Vousden, 2013).

Additionally, mutations can occur in the N-terminal transcriptional transactivation domain.

It is also common that the half-life of the mutant p53 is increased because of changes in

MDM2 interactions, and thereby mutant p53 becomes stabilized and expressed to a higher

level in tumor cells (Bartek et al., 1991). In breast cancers, mutated p53 can be found in

the cytoplasm of tumor cells, indicating a transcriptional inactivation of p53 or folding of

the p53 caused by R175H, structural mutants (Moll et al., 1992).

Majority of p53 functions are regulated through transcriptional regulation of target genes.

When p53 is induced by DNA damage, it blocks forward progress at the G1 phase and

thereby, p53 acts as “guardian of the genome” and drives a cell cycle “checkpoint”

function by transcribing. An important target gene of p53 is p21, which is an inhibitor of

CDK2 and Cdc2 that are active in the late G1, S, G2, and M phases of the cell cycle.

Furthermore, p53 is shown to transcribe several genes (Figure 3) when inducing cell cycle

arrest and when mediating DNA repair (reviewed in Riley et al., 2008). In addition, mouse

models of p53 knockouts show increased genomic instability and broader tumor spectrum

(Fridman and Lowe, 2003; Lowe et al., 2004). The p53 has been also suggested to regulate

apoptosis both transcription-independent manner and -dependent manner, where it directly

transcribes proteins like Puma and Bax (Speidel, 2010). Several oncogenes, like E1A,

Myc and E1F, induce p53 and therefore inactivation of p53 severely compromises

oncogene-induced apoptosis. Moreover, studies in mice indicate that disruption of p53-

mediated apoptosis co-operates with oncogenes in multiple mouse cancer models. In

Page 21: Apoptosis by c-Myc - Helsinki

addition, p53 can promote senescence in MEFs and senescence is a prerequisite for the

transformation of cells (Lowe and Sherr, 2003). Also, p53 transcribes for example TSP1,

which is a blocker of the development of new blood vessels and is shown to be inhibiting

angiogenesis. Emerging evidence also suggest that p53 regulates metabolic pathways such

as glycolysis and thereby enable stress adaptation (Berkers et al., 2013). Recently, more

transcription-independent functions of p53 has been emerging such as regulation of the

microRNA expression and maturation pathway and, also, direct role in DNA repair

process (Sengupta and Harris, 2005; Suzuki and Miyazono, 2013). Transcriptional

independent functions of p53 in apoptosis will be discussed more in 2.3.1.3.

Figure 3. The molecular target and effects of tumor suppressor p53. Modified from (Brown et al., 2009).

LKB1 (PAR-4) is a tumor suppressor protein and a serine/threonine kinase 11, lost in an

autosomal inherited Peutz-Jeghers syndrome (PJS) cancer patients (Jeghers et al., 1949;

Martin and St Johnston, 2003). Patients typically develop multiple benign gastrointestinal

polyps, hamartomas, and mucocutaneous pigmentation in different area and, they are at

risk for intestinal or extra-intestinal cancer (Martin and St Johnston, 2003; Vaahtomeri and

Makela, 2011). In addition, Lkb1 is one of the most commonly mutated genes in sporadic

human lung cancer (NSCLC) and, carriers of germline Lkb1 mutations have increased

Page 22: Apoptosis by c-Myc - Helsinki

breast cancer incidence (Hearle et al., 2006; Shackelford and Shaw, 2009). In tumors,

Lkb1 mutations are suggested to correlate with aggressive phenotypes of tumors and cause

increased invasiveness in mouse models (Vaahtomeri and Makela, 2011). Mammalian

Lkb1 has been implicated as a regulator of multiple biological processes: including

angiogenesis, cell cycle arrest, transformation (Wnt- and Ras pathways), p53-mediated

apoptosis, polarity, and energy metabolism (reviewed in Baas et al., 2004b). Lkb1 is

regulated via binding partners that anchors Lkb1 in the cytoplasm. Furthermore, the kinase

activity is crucial for the tumor suppressive function of Lkb1 (Ylikorkala et al., 1999).

Several groups identified Lkb1 as an upstream kinase and regulator of 5´AMP-activated

protein kinase (AMPK), which is a metabolic stress regulator. Thereby, Lkb1 has been

suggested to act as a tumor suppressor of metabolic transformation of cancer cells

(Shackelford and Shaw, 2009). In contrast, Lkb1 was shown to activate AMPK by

phosphorylating Thr172 in the regulatory activation loop, T-loop (Hawley et al., 1996).

The Lkb1 kinase activity is regulated via the Lkb1-STRAD-MO25 complex resulting in

activation of the AMPK and its related kinases (ARKs, 14 in total) (Baas et al., 2004b;

Vaahtomeri and Makela, 2011). In addition, Lkb1 inhibits proliferation and growth by

affecting mitogenic pathways such as mTOR (S6K1). Lkb1 promotes also senescence in a

p53-independent manner via Nuak2 and, in contrast, cell cycle arrest and cell death in a

p53–dependent manner (Humbert et al., 2010; Tiainen et al., 2002). Part of the Lkb1-

induced regulation of p53 is suggested to occur by direct Lkb1-mediated phosphorylation

of p53 Ser15, or its substrate kinases, AMPK, in response to glucose starvation or cell

detachment (Jones et al., 2005).

Lkb1 belongs to a family of partitioning-defective (PAR) protein that are required for

asymmetrical cell division of the C. elegans zygote. In addition, loss of Lkb1 was

discovered to disrupt polarization of the oocyte cytoskeleton resulting in a loss of anterior-

posterior cell polarity in Drosophila melanogaster and is required for apical-basal polarity

in epithelial cells (reviewed in Baas et al., 2004a). In epithelia, the cell-ECM interface

defines the basolateral membrane, whereas highly organized tight (TJ) and adherens

junctions (AJ) form a physical border between the apical and basolateral membrane

domains of a cell (reviewed in Baas et al., 2004b; Shin et al., 2006). In general, junctions

form a physical border between the apical (extending to the lumen) and basolateral

membrane domains of a cell provide strength to the epithelial sheet by serving as

Page 23: Apoptosis by c-Myc - Helsinki

anchoring sites for the cytoskeleton. Polarity is formed when apical microvillis are

assembled, TJs and AJs are formed, and finally the apical and basolateral surface markers

are ordered. Furthermore, cell-ECM contacts by integrins and cell-cell contacts by

cadherins regulate the transition of a nonpolarized to polarized epithelial cell transition

(Drubin and Nelson, 1996). Activation of LKB1 in mammalian epithelial cell lines results

in the cell´s intrinsic polarization via actin cytoskeleton reorganization and junctional

protein ZO-1 re-localization and involves STRAD (Baas et al., 2004a). The potential role

of Lkb1 in the regulation of polarity is supported by data that Lkb1 deletion in mouse

mammary glands leads to development of ductal carcinomas (McCarthy et al., 2009).

Several studies have linked the polarity functions of Lkb1 with its kinase activity and

downstream target AMPK, for example, AMPK null mutations lead to a strong polarity

phenotype with defects in Drosophila (Mirouse and Billaud, 2011). In contrast, in PJS

polyposis Lkb1 is thought to operate via TGF-β, and regulate cell polarity by controlling

microtubule and actin dynamic by Smad4, Nuak2 or p53 (Vaahtomeri and Makela, 2011).

AMPK is an evolutionarily conserved fuel-sensing serine/threonine protein kinase enzyme

that is activated under stress such as: hypoxia, ischemia, glucose deprivation, and exercise

(Steinberg and Kemp, 2009). Additionally, AMPK is suggested to be a metabolic tumor

suppressor (Figure 4), whose activity is often decreased in metabolic disorders like insulin

resistance, obesity, and Type 2 diabetes. Several epidemiological studies indicate that

metabolic syndrome increases the risk of cancer and thus, patients with Type 2 diabetes

taking well-accepted AMPK activator, metformin, have a reduced risk of cancer (Hardie,

2007; Luo et al., 2010). However, very little data exists on reduced AMPK activation in

cancer.. For example, AMPK mRNA levels inversely correlate with clinical prognosis in

breast and ovarian tumors and AMPK activation is reduced in both lung cancer specimen

containing loss-of function mutations of LKB1 and in breast cancer specimens (Conde et

al., 2007). In addition, a large number of studies have shown that activation of AMPK (by

pharmacological agents, exercise and dietary restriction) can attenuate cancer cell growth

in vitro and inhibit tumor development in vivo (Jiang et al., 2008). Conversely, as energy

stress sensor, AMPK activates catabolic pathways, which can promote tumorigenesis and

will be discussed later.

Page 24: Apoptosis by c-Myc - Helsinki

Figure 4. AMPK molecular targets and effects. Modified from (Liang and Mills, 2013).

AMPK is a heterotrimeric complex consisting of three subunits: a catalytic subunit (α),

and two regulatory subunits (β and γ) (reviewed in Hardie, 2007). Under metabolic stress

such as adequate carbon source and oxygen, the intracellular AMP levels or the AMP to

ATP ratio (ADP/ATP goes down) is increased and AMP is bound to the γ-subunit.

Subsequently, AMPK undergoes allosteric activation, thereby protecting it against

phosphatases that dephoshorylate threonine 172 in the activation loop (Hardie, 2007).

Several other AMPK activators are known, even though their exact mechanism is unclear

(cytokines, activating drugs: metformin, phenmorfin, AICAR, A-769662, and some

natural plant products) (Hardie, 2007). Thus far, several kinases have been identified to

phosphorylate threonine 172 on the catalytic subunit, leading to its activation. Previous

data suggests that in most scenarios, the LKB1 is kinase responsible for AMPK

phosphorylation (Shaw et al., 2004). However, AMPK is also a downstream substrate of

calmodulin-dependent protein kinase (CAMKK-β), which is activated due to increased

intracellular Ca2+ levels and TAK1 kinase (Hawley et al., 2005).

Page 25: Apoptosis by c-Myc - Helsinki

As a regulator of energy homeostasis (Figure 4) the activated AMPK inhibits lipogenesis.

It simultaneously stimulates fatty acid oxidation (ACC-2) and inhibits fatty acid synthesis

(ACC-1) in order to generate more ATP to cope with acute energy demands, and also

inhibits anabolic processes that consume ATP (Luo et al., 2010). In addition, AMPK

activation enhances insulin sensitivity, inhibits hepatic glucose production, stimulates

glucose uptake (by GLUT4 translocation) in muscle, and esterification alleviates

hyperglycemia and hyperlipidemia (Luo et al., 2010). The mechanisms behind these

functions involve down-regulated expression of biosynthetic genes involved in

gluconeogenesis and lipogenesis (Figure 4). Controversially, AMPK suppresses glycolysis

in tumor cells by inhibiting mTOR and on the other hand, it stimulates glycolysis by

activating PFK2 (Marsin et al., 2002).

Likewise, AMPK activation can regulate various processes (Figure 4), including cell cycle

checkpoint, cell polarity, senescence, autophagy, apoptosis, and pro-inflammatory

responses (Hardie, 2007; Shackelford and Shaw, 2009). Pharmacological activation of

AMPK is shown to inhibit growth of cancer cells by causing phosphorylation of p53 on

Ser15 and accumulation of p53 (Imamura et al., 2001). Also, glucose deprivation induced

arrest in MEFs is mediated by AMPK-mediated phosphorylation of p53 (Jones et al.,

2005). AMPK switches off protein synthesis by inhibiting elongation-factor-2 or

inhibiting the translation initiation step through rapamycin (TOR) pathway (Inoki et al.,

2003). Inhibition of the TOR pathway is also involved when AMPK is causing autophagy

(Inoki et al., 2003). Furthermore, at least a part of LKB1´s role in regulation of apico-basal

polarity in D. melanogaster undergoes through its downstream target AMPK (Lee et al.,

2007). Likewise, activation of AMPK induces repolarization of epithelial kidney and

intestinal epithelial cells by assembly of TJs (Hardie, 2007; Inoki et al., 2003). Also,

additional mechanisms for AMPK-induced polarity have been suggested recently, such as

the activation of Myosin II or, alternatively, through NUAK1 (Mirouse and Billaud, 2011).

Interestingly, recent studies indicate that AMPK plays a role in linking metabolic

syndrome and cancer. AMPK is an essential mediator of the tumor suppressor LKB

1 and could be suppressed in cancer cells containing loss-of-function mutations of LKB1

or in cancers associated with metabolic syndrome. Finally, AMPK may function as a

metabolic tumor suppressor regulating glucose, lipid, and protein metabolism.

Page 26: Apoptosis by c-Myc - Helsinki

1.2.3 Cancer metabolism

During tumor development, a growing tumor is soon running into nutrient and oxygen

lack and to bypass such metabolic stress and continue proliferation, tumor must undergo a

period of metabolic adaptation to survive or alternatively cells undergo apoptosis. The

altered metabolism found in cancer was discovered by Otto Warburg in 1920s, when he

described a phenomenon termed the ”Warburg effect”; an increase in glycolysis that is

maintained in conditions of high oxygen tension (“aerobic glycolysis”) and gives rise to

enhanced lactate production (Warburg et al., 1927). Now it is recognized that several

changes occur in the tumor cell´s metabolism, which result in a “metabolic

reprogramming” involving aerobic glycolysis, de novo lipid biosynthesis and glutamine-

dependent anaplerosis, glutaminolysis (Figure 5). These changes support growth and

allow necessary nutrients, energy (rapid ATP production,) and biosynthetic activity

(macromolecules) that are required for the increased proliferation of tumor cells

(DeBerardinis et al., 2008). Moreover, redox status is crucial for cancer cells, and low

levels of reactive oxygen species (ROS, byproducts of metabolic processes) can benefit

tumor cell´s proliferation for instance by activating Src kinase or inactivating PTEN.

However, at high levels ROS starts to induce damage and death. To counter this ROS-

induced oxidative stress, cancer cells also have a highly active antioxidant defense-system

still allowing, for example, moderate levels of ROS to induce mutagenesis (reviewed in

Cairns et al., 2011). Another important feature of cancer cells is increased nutrient uptake,

such as glucose uptake, that can be taken advantage of when imaging primary tumors and

metastases in clinics using the FDG-PET technique (Mankoff et al., 2007).

Several reasons explain why enhanced glucose uptake for glycolytic ATP generation or

anabolic reactions constitutes an advantage for tumor growth. Cancer cells rely on aerobic

glycolysis instead of oxidative phosphorylation (OXPHOS) for ATP production and

generate bicarbonic and lactic acids as end products of glycolysis (Pouyssegur et al., 2006).

Once glucose is transported inside of the cells it is converted to pyruvate that subsequently,

enters either the mitochondria for further conversions to acetyl-CoA, or is immediately

converted to lactate (reviewed in Cairns et al., 2011). In normal cells, glucose is converted

into acetyl-CoA and its complete oxidation occurs through the mitochondrion-localized

tricarboxylic acid (TCA) cycle. Furthermore, oxidative phosphorylation produces electron

donors NADH and FADH2, which donate electrons to the respiratory chain complexes I

Page 27: Apoptosis by c-Myc - Helsinki

and II, respectively CO2 and H2O as end products (total net of 38 ATP molecules). In

contrast, in tumor cells glucose is converted only to pyruvate (generates 2 ATP molecules),

and subsequently to waste products. These anaerobic waste components, acids, have their

own effect on the microenvironment like favoring tumor invasion, suppressing anticancer

immune effectors, and/or they are used by other cells (Kroemer and Pouyssegur, 2008).

Moreover, tumors can metabolize glucose through the pentose phosphate pathway (PPP)

to generate NADPH, which is required for macromolecule biosynthesis and also

contributes to fatty acid synthesis. This is also a way in which cancer cells obtain

antioxidant defenses against a hostile microenvironment and chemotherapeutic agents

(Gatenby and Gillies, 2004). Furthermore, tumor cells can express the PKM2 isoform of

pyruvate kinase, than can surprisingly inhibit glycolysis or slow it down, thereby

promoting shuttling of pyruvate to the PPP instead (Schulze and Harris, 2012). Moreover,

cancer cells use or deviate intermediates of the glycolytic pathway for anabolic reaction

(cataplerosis), which in part enables increased growth (Figure 5).

In addition, pyruvate may enter a truncated TCA cycle resulting an exportation of citrate

from the mitochondrial matrix (Figure 5). Subsequently, its cleavage products acetyl-CoA

becomes available for the synthesis of fatty acid and cholesterol and furthermore,

oxaloacetate (OAA) is converted to malate and reimported to mitochondria into a citrate

cycle (reviewed in Kroemer and Pouyssegur, 2008). Importantly, it is speculated that the

reason why tumor cells use glycolysis is that it benefits both bioenergetics and

biosynthesis and if the glycolytic rate is high enough, a similar yield of ATP is produced

as in oxidative phosphorylation. Lactate is not used in the Warburg effect, and it starts to

accumulate in the cells. However, even though more pyruvate could be oxidized, it is not

done in cancer cell and instead, the cells need to get rid of excess pyruvate via a high-flux

mechanism, and pyruvate is converted to lactate by lactate dehydrogenase A (LDH-A)

(DeBerardinis et al., 2008). Furthermore, in glutaminolysis glutamine is oxidized in the

TCA cycle, allowing proliferating cells to use TCA cycle intermediates as precursors for

biosynthesis (Yuneva et al., 2007).

Page 28: Apoptosis by c-Myc - Helsinki

Figure 5. Metabolic alterations in cancer. Metabolic reprogramming of cancer cells involves activation catabolic pathways (b-oxidation, glycolysis, glutaminolysis), which release carbon for bioenergetic reactions and promotes biosynthesis of nucleic acids. In addition, biosynthesis requires enhanced nutrient uptake. Part of the biosynthesis required for cell growth comes from cataplerosis, where metabolic intermediates are directed for synthesis, for example, glucose 6-phosphate of glycogen and ribose 5-phosphate synthesis, dihydroxyacetone phosphate for triglyceride and phospholipid synthesis, and pyruvate for alanine and malate synthesis. Caner cell have Truncated TCA cycle, in which citrate is exported into the cytosol and used for lipogenesis. Tumor cells also produce ATP via oxidative phosphorylation (OXPHOS), which imposes an additional metabolic burden on the TCA cycle. Modified from (Galluzzi et al., 2013). In general, the metabolism in cancer cells is regulated autonomously without the

requirement of growth-factor-induced signals. Variety of molecular changes explains this

cancer-specific “metabolic reprogramming”, such as mitochondrial DNA mutations,

oncogene activation, mutation in the Ras and PI3K/Akt/mTOR pathway and aberrant

activation of the AMPK-LKB1 pathway (Cairns et al., 2011). For instance, constitutive

cellular stabilization of (hypoxic-) stress- and oncogene-induced HIF-1 is one major

reason for aerobic glycolysis as decreasing conversion of pyruvate to acetyl-CoA, and this

compromises OXPHOS (Harris, 2002). In contrast, tumor suppressor p53 maintains the

respiration, OXPHOS, by increasing the expression of cytochrome c oxidase 2 and down-

regulates phosphoglycerate mutase. Moreover, p53 transcriptionally activate hexokinase-2

Page 29: Apoptosis by c-Myc - Helsinki

and TIGAR, which inhibits phosphofructokinase activity, lowers the levels of FBP, and

hence inhibits glycolysis while channeling glucose to the PPP (Bensaad et al., 2006). In

addition to the genetic changes that alter tumor cell metabolism, the abnormal tumor

microenvironment such as hypoxia, pH and low glucose concentrations have a major role

in determining the metabolic phenotype of tumor cells (Cairns et al., 2011). Another

structural feature of tumor cells is that mitochondria are often small, lack cristae, and are

deficient in the F1 subunit of ATP-synthetase. However, there is no evidence that

mitochondrial respiration would be functionally impaired (Kim and Dang, 2006; Kroemer

and Pouyssegur, 2008). Interestingly, the resistance of cancer cells mitochondria to

apoptosis-associated permeabilization and subsequent cytochrome c release, and the

altered contribution of these organelles to metabolism, are closely related. For example,

increased glycolysis in tumor cells is accompanied by increased stability of the

mitochondrial membrane partly by membrane-associated hexokinase (HK) I/II

upregulation in tumors. Furthermore, defects in OXPHOS induce apoptosis resistance

(Tomiyama et al., 2006). Other alterations in cancer and disabled apoptosis are

hyperpolarization of the inner mitochondrial transmembrane potential and deficiency in

voltage-gated plasma membrane K+ channels (Kroemer et al., 2007).

Mammary gland is an organ that produces milk in the female breast and consists of

epithelial cells lining the walls of alveoli, where the milk is secreted, and surrounding

myoepithelial cells. The alveoli join to form groups known as lobules ending in the nipple.

Beneath the epithelial cell layers lies a basement membrane, which separates the epithelial

cells from the underlying layer of connective tissue cells, the stroma. Furthermore, the

basement membrane consists of extracellular matrix (ECM), which together with

adipocute, fibroblast, inflammatory cells constitute mammary stroma (Watson and Khaled,

2008). Importantly, breast cancer affects one in eight women during their lifetime. Breast

cancer can be classified by different schemata, which reflect treatment responses and

prognoses. Categories include histopathological type, grade (1; well-differentiated and

best prognosis, 2; moderately-differentiated and medium prognosis, 3; poorly-

differentiated and worst prognosis), stage (tumor-node-metastasis; TNM), receptor status

and the presence or absence of genes as determined by DNA testing. The majority of

Page 30: Apoptosis by c-Myc - Helsinki

breast cancers derive from epithelia and are classified as mammary ductal carcinoma

(Pinder, 2010). Ductal carcinoma in situ (DCIS, 13%) is proliferation of cancer cells

within the epithelial tissue without invasion of the surrounding tissue and is harmless.

However, 60% of the lesions will become invasive over the course of 40 years in follow-

up (Zorbas et al., 2004). In contrast, invasive ductal carcinoma (55%) and invasive lobular

carcinoma (5-10%) invades the surrounding tissues. Treatment options always include

surgery for removal of the main tumor mass, or mastectomy (removal of the breast) and

adjuvant therapy includes chemotherapy, radiotherapy, hormonal therapy, and/or targeted

therapy.

The receptor status of the breast cancer is identified by immunohistochemistry, and it

defines the presence of estrogen receptor (ER), progesterone receptor (PR), and HER2. In

addition, tumors also exhibit different gene expression profiles and based on mRNA

profiling data, breast cancers have been classified into 5 subgroups: luminal A and luminal

B, HER2 amplified, basal-like, and normal-like (Perou et al., 2000). Of these, tumors of

the luminal type are generally ER/PR positive, whereas the basal-like and normal-like

subtypes are triple-negative (negative for ER, PR and HER2 receptors). In addition to lack

of receptors, the characteristics of the basal-like subtype are high expression of basal

markers such as cytokeratin 5/6 and/or increased expression of EGFR. Receptor status is a

critical assessment for all breast cancers as it determines the suitability of using targeted

treatments. ER positive cancer cells depend on estrogen for their growth, so they can be

treated with drugs to reduce either the effect of estrogen (e.g. tamoxifen) or the actual

level of estrogen (e.g. aromatase inhibitors). HER2 positive cells respond to drugs such as

the monoclonal antibody, trastuzumab (Gonzalez-Angulo et al., 2007). Conversely, triple

negative (ER-, PR-, HER2) cancer, lacking targeted treatments has a poorest prognosis

(Oakman et al., 2010). Additionally, DNA microarrays can be used for further

categorizing of breast cancer subtypes, such as basal-like.

Furthermore, Myc deregulation contributes to breast cancer initiation and progression and

is associated with poor outcomes (Robanus-Maandag et al., 2003). Myc is amplified

approximately in 16 % of all breast cancers and overexpressed 20 45% of cases,

depending on the study (Bieche et al., 1999; Chrzan et al., 2001; Naidu et al., 2002). The

basal-like breast cancer constitutes one of the most challenging subtypes of breast cancers

Page 31: Apoptosis by c-Myc - Helsinki

accounting for 15% of breast cancer cases. In addition, the basal-like breast cancer is

responsible for most breast cancer deaths, since they lack adequate hormonal therapy.

Importantly, high Myc pathway activity has been implicated in basal-like breast tumor

subtype based on pathway analyses from molecular profiling (Xu et al., 2010).

Cell death, or “cell suicide”, is essential for proper tissue homeostasis and adaptation to a

changing environment and the importance of this phenomenon is emphasized by the fact

that the components of cell death are well conserved (Adams 2003). There are many

triggers for cell death, such as physical or chemical stress, parasites or developmental

signals (embryogenesis and metamorphosis). Originally, Kerr, Wyllie and Currie were the

first to introduce the term “apoptosis” as a form of programmed cell death, to describe the

morphological process leading to controlled cellular self-destruction (Kerr et al., 1972).

Importantly, dysfunction or deregulation of the apoptotic program is implicated in a

variety of pathological conditions like cancer, autoimmune diseases, and the spreading of

viral infections, while neurodegenerative disorders, AIDS, and ischaemic diseases are

caused or enhanced by excessive apoptosis. Currently, cell death types involve: apoptosis,

autophagic cell death, necrosis, necroptosis and pyroptosis, which all are characterized by

a number of distinguishing features and morphological changes (Degterev and Yuan,

2008). First of all, in apoptosis the contents of the dying cell remain contained in

membranes. Another feature is that the cell shrinks, shows deformation and detaches from

surrounding cells, (loss of adhesion), the plasma membrane shows blebbing, and finally,

the cell fragments into smaller, membrane-bound structures known as “apoptotic bodies”.

In addition, the nucleus undergoes characteristic changes, including chromatin

condensationand internucleosomal cleavage of DNA leading to an oligonucleosomal

“ladder” appearing in agarose gel (Wyllie et al., 1980). One biochemical feature is

proteolytic cleavage of a number of intracellular substrates, however, organelle integrity is

still maintained. Finally, the apoptotic bodies are phagocytosed in a relatively short time

by macrophages and parenchymal cells, which recognize the externalized PS on the outer

membrane (reviewed in Saraste and Pulkki, 2000). Thus, in most of the cases, apoptotic

Page 32: Apoptosis by c-Myc - Helsinki

bodies are removed from the tissue without a subsequent inflammatory response, however,

if not phagocytosed, the cell will undergo degradation resembling necrosis.

Autophagic cell death (macroautophagy) differs from apoptosis and is less well

understood. It involves a cellular mechanism of self-eating, autophagy, which can finally

lead to cell death. However, autophagy can also be involved in the turnover of long-lived

proteins and whole organelles (mitophagy or reticulophagy), or in the cell´s adaptation to

starvation-induced stress and lack of nutrients (Maiuri et al., 2007). During autophagy, the

cytoplasm is engulfed by specific double- or multi-membraned autophagic vacuoles

(autophagosomes) and later digested by lysosomal enzymes. Importantly, apoptosis and

autophagy are sometimes overlapping and can serve as an alternative choice if the other

one is inhibited. The third type of death is necrosis, in which cells are suffering a major

insult, resulting in a loss of membrane integrity, swelling and disruption of the cells

leading to a release of cellular contents. Thus causing damage to surrounding cells, and a

strong inflammatory response in the corresponding tissue. However, there are forms of

necrosis that are programmed, termed “necroptosis” (Golstein and Kroemer, 2007). In

addition, “pyroptosis “ is a programmed cell death form associated with antimicrobial

responses during inflammation (Fink and Cookson, 2005).

Apoptosis is mediated by signaling pathways, the mitochondrial (intrinsic) and death

receptor apoptotic (extrinsic) pathway, finally leading to disruption of the cell (Fulda and

Debatin, 2006). In addition, this process is regulated by two main groups of proteins;

caspases and the Bcl-2 family. The history of finding key regulators of apoptosis and

vertebrate homologs began with studies on C. elegans and its molecular machinery

required for proper development: proapoptotic caspase CED-3 and adaptor protein CED-4

and anti-apoptotic CED-9 (Bcl-2 homolog) (Ellis and Horvitz, 1986). Currently it is

appreciated that the balance between the Bcl-2 family interactions regulates the

mitochondrial apoptosis pathway, which upon activation further results in caspase

activation. Moreover, caspases can be catalytically activated to initiate a caspase cascade,

leading to degradation of the cellular proteins.

Page 33: Apoptosis by c-Myc - Helsinki

2.2.1 Caspases as executioners Caspases are cysteine-dependent aspartate-specific proteases residing in inactive form as

zymogens. Once activated, these endopeptidases recognize cysteine-residue in peptide

motif, QACRG, within a proteolytic substrate and cut that protein internally after the

aspartic acid residues (Stennicke and Salvesen, 1997). Since finding the human homolog

of C. elegans´s CED-3, the ICE (interleukin-1b-converting enzyme), 14 distinct

mammalian caspases has been identified in total, of which there are 12 in humans.

Subsequently, the family of apoptotic caspases has been divided into two major classes:

the initiator caspases including procaspase-2, -8, -9 and -10 (Dronc and Dredd in D.

melanogaster); and the effector (or executioner) caspases, which include procaspase-3, -6,

and -7 (Drice, Decay, Damm, Dcp1 and Stripa in D. melanogaster) (reviewed in Degterev

et al., 2003; Hay and Guo, 2006; Riedl and Shi, 2004). The ICE family contains caspase-

13, -5, -4, and -1, which mostly mediate cytokine maturation during inflammation,

however caspase-1 also mediate pyroptosis (Bergsbaken et al., 2009).

Inactive procaspases exists as a monomer carrying at its N-terminus the pro-domain,

followed by a large and a small subunit, which are sometimes separated by a linker

peptide (Figure 6). Activation of initiator caspases occurs at signaling complexes via pro-

domains that are protein-protein interaction motifs, whereas effector caspases are activated

by initiator caspases. The maturation of initiator procaspases at signaling complexes

involves dimerization, autoactivation, and proteolytical processing between the large and

small subunit, resulting in a separation of these subunits (reviewed in Riedl and Shi, 2004).

The recruitment to signaling complexes can happen either in response to the ligation of

cell surface death receptors or in response to signals originating from inside the cell.

Moreover, the pro-domains are mediating the interaction between caspases and adaptor

proteins, and caspase homodimerization. In the case of procaspase-8 and -10, the pro-

domain is called the death effector domain (DED) and in the case of procaspase-9 and -2

the caspase recruitment domain (CARD) (Figure 6). DEDs and CARDs are not related by

sequence homology, but they are structurally similar, each having a “death fold” structure.

Furthermore, caspase-8 and -10 are recruited to the death-inducing signaling complex

(DISC), caspase-2 to PIDDosome and caspase-9 to apoptosome (Boldin et al., 1995; Tinel

and Tschopp, 2004; Zou et al., 1999).

Page 34: Apoptosis by c-Myc - Helsinki

Figure 6. Structure and classification of mammalian caspases. Caspases can be classified on the basis of their function in apoptosis, inflammation or differentiation. All caspases are characterized by a catalytic domain (p20 and p10 subunit) that can give rise to large and small subunits upon processing, and an N-terminal prodomain of variable length. The long prodomain caspases contain specific protein-protein interaction domains (CARD, DED) that are involved in the recruitment of caspases in specific multiprotein caspase-activation platforms (DISC, inflammasome, apoptosome, PIDDosome). caspase-12 is expressed as a truncated, catalytically inactive protein in most humans. Modified from (Riedl and Shi, 2004)

According to the “proximity-induced dimerization model”, when several procaspase-8

molecules are bound, for example, to DISC, they are in close proximity to each other and

are therefore assumed to activate and cleave each other by autoproteolysis (reviewed in

Denault and Salvesen, 2002). However, the activation of initiator caspases is still partly

unknown. When active, for example in the case of caspase-8, the heterotetramer consists

of two small and two large subunits, whereas caspase-9 exists predominantly as a

monomer (Degterev et al., 2003). More importantly, the cleavage does not directly

activate initiator caspases instead it brings a specificity-binding pocket into the correct

position, and thereby stabilizing the caspases. Also, other proteins with sequence

similarity, like FLIP and MALT, can dimerize with caspase-8 and form an active site.

However, without a proper proteolytic activity this does not results in caspase-8 cleavage

and apoptosis is inhibited. Furthermore, the FLIP-caspase-8 heterocomplex has been

suggested to mediate non-apoptotic signaling pathway that does not require processing of

caspase-8 (Micheau et al., 2002).

Page 35: Apoptosis by c-Myc - Helsinki

In contrast, effector (executioner) caspases constitutively pre-exist as inactive homodimers

and possess only short pro-domains, therefore lacking the ability to auto-activate (Figure

6). Thus, the activation of effector caspases is carried out by initiator caspases via

cleavage between the large and small subunits (reviewed in Riedl and Shi, 2004). Effector

caspases are responsible for the final step in apoptosis, and many of their target proteins of

them participate in the formation and regulation of the cytoskeleton, function at the cell-

to-cell and cell-to-matrix attachment points or are involved in the regulation of chromatin

structure (Riedl and Shi, 2004). For example, an apoptotic cell undergoes extensive

plasma membrane blebbing, where caspases cut actin regulator proteins resulting in actin

polymerization and, finally, blebbing. Also, “eat met” signals, lipid phosphatidylserine

externalization, of apoptotic cells are also produced by caspase activation (Fischer et al.,

2003). An important caspase-3-mediated step is chromatin fragmentation, where caspase-3

releases nucleases (caspase-activated DNase) by cutting their inhibitor protein, (iCAD)

(Enari et al., 1998). Another final step is chromatin condensation, which is caused by

caspase-3 and -7-mediated cleavage of the acinus protein, and the disruption of the nuclear

envelope resulting from caspase-6-mediated cutting of lamins (Nicholson, 1999). In

addition, caspases have an effect on ATP levels and energy homeostasis since they cut the

mitochondrial electron transport protein, NDUFS1, (Ricci et al., 2004a).

Furthermore, the enzymatic activity of caspases is inhibited by the IAP family of proteins

(XIAP, c-IAP1 and c-IAP2) (Deveraux et al., 1998). In Drosophila DIAP1 is one of the

most important inhibitor of apoptosis and binds DRONC. The cIAPs contain baculovirus

IAP repeat BIR domains 1-3, through which they, for example, compete to bind to active

binding pocket (Vaux and Silke, 2005). In addition, many IAP proteins contain another

domain called RING, that functions as an E3-ligase, responsible for putting ubiquitin

chains on target proteins and targeting them for degradation. During apoptosis, the IAP-

mediated inhibition of caspases is antagonized by another family of proteins (such as

SMAC/Diablo), which contain an IAP-binding tetrapeptide motif (reviewed in Vaux and

Silke, 2005).

Page 36: Apoptosis by c-Myc - Helsinki

2.2.2 Bcl-2 family The Bcl-2 family of proteins were named for the first family member to be described, Bcl-

2, which was discovered as an overexpressed oncogene in B-cell follicular lymphoma

(Tsujimoto et al., 1984). Currently, the Bcl-2 family consists of 12 proteins in mammals,

which regulate the mitochondrial apoptosis pathway, a process called mitochondria outer

membrane permeabilization (MOMP), and cytochrome c release from mitochondria

(Figure 7). The Bcl-2 family is defined by the presence of conserved motifs known as Bcl-

2 homolog (BH) domains (BH1 to BH4), which promotes the interaction and dimerization

within the family (reviewed in Chipuk et al., 2010). Cellular stresses cause transcriptional

and posttranscriptional regulation of the Bcl-2 family resulting in MOMP. According to

the BH domains structure and function, the family of Bcl-2 proteins is divided into pro-

and anti-apoptotic members (Figure 7). Additionally, Bcl-2 family protein may have

alternate functions in other homeostatic pathways including glucose metabolism, cell

cycle and regulation of mitochondrial morphology (Danial, 2007).

The pro-apoptotic Bcl-2 effectors Bak and Bax, (and Bok), contain multiple domains and

are responsible for MOMP, pore forming and the actual cytochrome c release. Once

activated, Bak and Bax bury themselves in the mitochondrial outer membrane, where they

form oligomers of multiple sizes and are suggested create holes (not channels or pores),

which allow large molecules to pass through the membrane (Youle and Strasser, 2008).

However, visualization of these oligomers has shown that large clusters of Bax are

localized near, but not on the OMM, so pore formation is controversial. Bax is shown to

be soluble in the cytosol and during apoptosis it translocates to mitochondria and is

inserted into OMM (Bleicken et al., 2010; Hsu et al., 1997). In contrast, Bak is suggested

to localize at OMM all the time in the cells by its carboxy-terminal region. More precisely,

in viable cells Bak and Bax exist as monomers or dimers and are said to be inactive as

measured by conformation with N-termini being hidden (Letai et al., 2002). Furthermore,

the presence of Bak can facilitate opimal Bax translocation (Mikhailov et al., 2003). Mice

lacking either one, Bak or Bax develop normally and apoptosis is efficient, however,

animals lacking both Bax and Bak are lethal to the embryo. This indicates a molecular

redundancy in which either Bax or Bak can effect the MOMP (Wei et al., 2001).

Page 37: Apoptosis by c-Myc - Helsinki

Figure 7. Classification of Bcl-2 family proteins. BH: Bcl-2 homology domain; TM: transmembrane domain. Modified from (Taylor et al., 2008).

Another class is the proapoptotic BH3-only proteins (Bad, Bid, Bik, Bim, Bmf, Hrk, Noxa,

and Puma), which share only a conserved α-helical BH3 domain through which they

regulate other members. The different BH3-only proteins have varied tissue distributions

and are expressed under different conditions. Furthermore, they are targets of signal

transduction pathways and are therefore considered as sensors for environmental stress.

Bid is a sensor for protease and is unavailable for interaction with the Bcl-2 family

proteins when native. It has a large, flexible loop that can be cleaved by proteases like the

caspases and can subsequently insert into mitochondrial membranes; the BH3 domain

becomes exposed and can interact with other Bcl-2 membranes (Luo et al., 1998). In

contrast, Bim and Bad are sensors for growth factor deprivation (Lomonosova and

Page 38: Apoptosis by c-Myc - Helsinki

Chinnadurai, 2008). For instance, Bad is phosphorylated in response to survival factor,

which results in interaction changes with Bcl-xL and apoptotic resistance. (Zha et al.,

1996). Other BH3 only members have shown to interact with non-mitochondrial proteins,

for instance Bim is binding to the dynein light chain 1, and BMF associates with dynein

light chain 2 (Huang and Strasser, 2000). Additionally, Puma and Noxa are under

trrancriptional regulation by p53 and have important role during DNA damage-induced

apoptosis. The function of BH3-only protein will be discussed more at 2.3.1.2.

Anti-apoptotic (pro-survival) Bcl-2 proteins class includes Bcl-2, Bcl-xL, Bcl-w, Mcl-1

and A1 and all except Mcl-1 possess domains BH1-4 forming a hydrophobic pocket

(Figure 7-8) to which the BH3 amphipathic α-helix of the other members can bind (Tait

and Green, 2010). They can be integrated within the OMM but may also be localized to

the cytosol or ER to inhibit MOMP. Anti-apoptotic members interact or homodimerize

with proapoptotic proteins in order to inhibit their function (Chipuk et al., 2010).

Multidomain Bcl-2 family proteins all have binding pocket present tat is critical for their

functions (Youle and Strasser, 2008). This hydrophobic binding pocket consists of BH1

(portions of α4–α5), BH2 (α7–α8), and BH3 (α2) domains, which all have certain

amphipathic a-helices (Chipuk et al., 2010). Furthermore, anti-apoptotic Bcl-2 proteins are

suggested to inhibit oligomerization of Bak and Bax by binding to their BH3 regions.

When proapoptotic BH3-only proteins activate multidomain Bak/Bax their hydrophobic

face of BH3 helix is exposed. This enables BH4-only protein to interact with the

hydrophobic groove of multidomain dimerization partners. Furthermore, this is thought to

induce series of conformational changes of Bak/Bax: (1) α1/N-terminal exposure, (2)

transient BH3 exposure, (3) protection of the membrane-embedded groove, and (4)

increased proximity of embedded monomers (Chipuk et al., 2010). For example, the BH3

domain of Bak binds to a groove in Bcl-xL formed from a-helices containing BH1-3

domains (Figure 8). Once Bak is activated, it exposes its BH3 domain, which opens a

groove into which the BH3 domain on another Bak molecule can bind, forming a dimer

(Dewson et al., 2009). Bcl-xL, then, may prevent this Bak-Bak interactions and

oligomerization by binding to this exposed BH3 domain. Furthermore, Bcl-2 family

Page 39: Apoptosis by c-Myc - Helsinki

members have binding specificities and, for instance, Mcl-1 and Bcl-xL are major binders

to the Bak BH3 domain, whereas Bcl-2 and Bcl-xL bind well to BH3 of Bax and poorly to

Mcl-1. In addition, Bcl-2-Bax heterodimerization requires also the BH1-3 groove and a

novel interface involving the BH4 motif of Bcl-2 and of Bax containing the BH3 motif

(Ding et al., 2010). Furthermore, the BH3-only member binding profiles varies. Bid and

Bim are interacting with all anti-apoptotic members and also with Bak and Bax, whereas

Bad (to Bcl-2, Bcl-xL, Bcl-w) and Puma (to all) are suggested to bind mainly anti-

apoptotic proteins. In addition, Noxa (to Mcl-1) and Hrk (to Bcl-xL) seem to be very

specific in binding (Chipuk et al., 2010).

Figure 8. BH3:groove interactions in the regulation of Bak activation and oligomerization. According to “direct activation model” (discussed in 2.3.1.2) anti-apoptotic Bcl-2 family members (Bcl-xL) inhibit the activator BH3-only members (Bid) and the multidomain pro-apoptotic Bak/Bax by sequesterating and occupying their BH3 domains. Sensitized BH3-only proteins (Puma) can displace Bid from Bcl-xL and allow Bid to activate (N-terminal and BH3 exposure) and oligomerize Bak.

Page 40: Apoptosis by c-Myc - Helsinki

2.3.1 Mitochondrial apoptosis pathway The mitochondrial (intrinsic) pathway of cell death is engaged by various types of

intracellular stresses including growth factor withdrawal, DNA damage by ultraviolet

(UV) and γ-irradiation stress, cytoskeletal disruption, unfolding stresses in the

endoplasmic reticulum or accumulation of unfolded proteins or hypoxia. Furthermore,

mitochondrial pathway is engaged by death receptor apoptosis pathway and thereby death

signal is amplified as discussed in 2.3.2.2. (Fulda and Debatin, 2006). The key players in

mitochondrial pathway are the Bcl-2 family proteins that respond to apoptotic signals.

During apoptosis, the mitochondria release a number of factors from their intermembrane

space, like cytochrome c, Smac/Diablo, and AIF, which promote and amplify the

apoptotic cascade. More recently, the mitochondrial fragmentation has been shown to be

important early stage during apoptosis and regulated by proteins such as Drp-1 and Fis1

(Karbowski et al., 2006).

Once the mitochondrial apoptosis pathway becomes activated by upstream signals, the

mitochondrial outer membrane permeabilization (MOMP) event occurs. This step is

sudden and rapid, and is considered to be the ‘point of no return’ during apoptosis (Tait

and Green, 2010). MOMP results in the diffusion of soluble molecules (cytochrome c)

residing in the intermembrane space to the cytosol. Of note, MOMP does not involve a

loss of integrity of the inner mitochondrial membrane and therefore, mitochondrial

function is not destroyed completely (Ow et al., 2008). Among the released proteins is

Smac/Diablo that binds same region of XIAP that binds to initiator caspase-9 and thereby

prevents XIAP from inhibiting effector caspases (Deveraux et al., 1997; Verhagen et al.,

2000). In addition, another protein released following MOMP is Omi (HtrA2), which,

like Smac, has an amino-terminal sequence that inhibits XIAP (Verhagen et al., 2002). In

addition, other lethal proteins released by MOMP are endoG, which is a mitochondrial

enzyme that cleave DNA between nucleosomes and AIF, which translocates from the

mitochondria to the nucleus and causes chromatin condensation and DNA fragmentation

(Susin et al., 1999). Importantly, the release of cytochrome c during MOMP results in

Page 41: Apoptosis by c-Myc - Helsinki

caspase activation required for cell death. Moreover, APAF1 is a CARD domain-

containing adaptor protein for initiator caspase-9 (Figure 6) and pre-exists in the cell as a

cytosolic, inactive monomer that cannot bind or dimerize caspase (Bratton and Salvesen,

2010). Whereas, the cytochrome c is a nuclear-encoded protein that is synthesized as apo-

cytochrome c and is transported into the mitochondria to the space between the inner and

outer mitochondrial membranes (the intermembrane space), where it becomes a

holocytochrome c (Ow et al., 2008). Upon cytochrome c´s release, it engages APAF1 and

changes its conformation, leading to exposure of the oligomerization and CARD domain.

Subsequently, the center of the APAF1 oligomer (CARDs) recruits caspase-9 to activate it

(Figure 10); this APAF1-caspase-9 complex is called the “apoptosome” (Reubold and

Eschenburg, 2012). This function of cytochrome c is suggested to be independent of its

function in electron transport, and is required for proper caspase activation during the

mitochondrial apoptosis pathway.

Mitochondrial permeability transition (MPT) occurs in the mitochondrial inner membrane

by a channel called the permeability transition pore (PTP). Exposure to high

concentrations of calcium or signals including reactive oxygen species, changes in cellular

pH and certain drugs can result in MPT. During MPT the voltage potential changes and

solutes enter the matrix and water swells the mitochondria until the inner membrane

ruptures the outer membrane. In addition, PTP is mostly composed of adenosine

nucleotide transporter (ANT) together with voltage-dependent anion channel (VDAC)

(Kroemer et al., 2007). Even originally thought, it is unlikely that MPT is a major

mechanism of MOMP in apoptosis.

Bcl-2 mediated controof MOMP is the major decision that determines whether a cell will

die by engaging the mitochondrial pathway of apoptosis. All of these decisions depend on

interactions among members of the Bcl-2 family of proteins. According to the first model

described, a “rheostatit model”, the balance between anti- and proapoptotic Bcl-2 proteins

defines the apoptotic state of the cell. Later on, two competing models have been

suggested for Bak/Bax activation. According to a “neutralization/indirect model” (Figure

9) BH3-only would primarily bind anti-apoptotic proteins and thereby displace or

Page 42: Apoptosis by c-Myc - Helsinki

neutralize them resulting in Bak/Bax activation and apoptosis (Uren et al., 2007; Willis et

al., 2007). Moreover, one class including Bid, Bim and Puma bind all anti-apoptotic

members and are more potent in inducing apoptosis. Whereas another group, including

Noxa and Bad, engages only selected group of anti-apoptotic proteins (Chen et al., 2005).

For instance, Bak has been shown to be sequestered by Mcl-1 and Bcl-xL until displaced

by BH3 only proteins Noxa and Bad, leading to self-association of Bax/Bak molecules

and MOMP (Willis et al., 2005).

Figure 9. Indirect and Direct activation models for Bak/Bax activation. In the indirect activation model, antiapoptotic proteins bind and inhibit proapoptotic Bax and Bak. Whereas, BH3-only molecules bind antiapoptotic protein to displace them from Bax or Bak, which subsequently can undergo conformational change and oligomerize. In the direct activation model, BH3-only proteins are divided into activator and sensitizer groups. Antiapoptotic Bcl-2 proteins bind and sequester both the activator and sensitizer subclasses. Sensitizer BH3-only proteins occupy the binding pocket of antiapoptotic molecules, allowing activator BH3-only proteins to engage Bax/Bak (Danial, 2007).

However, often neutralization is not enough and additional BH3 activator is required.

Thus, according to a more recent “a direct activator/de-repressor model” (Figure 9) these

Page 43: Apoptosis by c-Myc - Helsinki

proteins are called “direct activators”. For instance, Bid and Bim and their BH3 peptide

can trigger Bax and Bak oligomerization, and permeabilization of synthetic membranes of

isolated mitochondria (Kuwana et al., 2002; Letai et al., 2002; Wei et al., 2000).

Furthermore, this BH3 peptide-induced Bak/Bax activation seems to occur in a stepwise

manner involving a subset of BH3-only proteins (Kim et al., 2009). Puma has also been

shown to promote Bak/Bax activation, however, whether this is via indirect or direct

mechanisms remains controversial (Chipuk et al., 2005). Moreover, anti-apoptotic proteins

can sequester also these direct activators and another BH3-only proteins (“sensitizer or de-

repressor”) can free the direct activator BH3-only protein to activate Bak and Bax (Figure

9). Recently, novel model have been suggested termed “priming-capture-displacement”

model (Strasser et al., 2011). Additionally, other proteins like p53 can exert an ability to

activate Bak and Bax (Kuwana et al., 2005). Importantly, it is suggested that direct

activators do not remain associated with the pro-apoptotic effector protein; instead this

interaction, for example between Bax and Bid, would be physical and transient. According

to the “embedded together model”, the critical step in Bax activation is the initial

membrane insertion and the model emphasizes interaction among Bcl-2 family members

and different conformations (Leber et al., 2007). However, how higher order oligomers

and pores are formed is still controversial. Furthermore, studies using BH3 peptides with

covalently stapled α-helices show that the BH3 domain of Bim binds to the “back” of Bax

and not to the same spot as Bcl-2 (Gavathiotis et al., 2008). Subsequently, Bax undergoes

a conformational change and exposes its BH3 and BH-binding groove and dimerizes.

Tumor suppressor p53 has an important role in regulating the mitochondrial apoptosis

pathway. Particularly, p53 is activated in response to a variety of DNA-damaging agents,

including IR and UV (Kastan et al., 1991). DNA damage results in an accumulation of

p53 protein through post-translational mechanisms, and subsequent increase in p53

activity results in cell cycle arrest or apoptosis (Zilfou and Lowe, 2009). In the nucleus,

p53 can mediate apoptosis by transcriptional activation of pro-apoptotic genes like the

BH3-only proteins for instance (Figure 3) and by transcriptional repression of Bcl-2 and

the IAPs (reviewed in Benchimol, 2001). However, induction of these target gene

products show variable kinetics, with some being delayed. Therefore, evidence for

Page 44: Apoptosis by c-Myc - Helsinki

transcription-independent p53-mediated apoptosis has been accumulating (Mihara et al.,

2003). It has been shown that p53 can directly translocate to the mitochondria after DNA

damage without transactivation domain, and in another analysis this precedes target gene

activation (Erster et al., 2004). In interaction analysis, p53 is suggested to bind Bcl-2 and

Bcl-xL. Furthermore, the p53-Bcl-xL interaction occurs either via the DNA binding domain

of p53 interacting with the α1/BH4 and partial α2/BH3 of Bcl-xL or via the N-terminus of

p53 interacting with Bcl-xL (Petros et al., 2004; Xu et al., 2006). More importantly, p53

interacts with amino acid residues distinct from those that mediate the binding of Bcl-xL to

Bak. In addition, naturally occurring mutations in p53 decrease the p53-Bcl-xL binding,

whereas mutated p53 is still able to bind, for example, Bak but further oligomerization is

abrogated (Mihara et al., 2003; Pietsch et al., 2008). First, p53 was shown to regulate

MOMP by direct activation of Bax, however, later p53 was demonstrated to interact via

the tetramerization domain with Bak and induce oliogmerization of Bak and disruption of

Bak-Mcl-1 complex (Chipuk et al., 2004; Pietsch et al., 2007). Altogether two possible

models are proposed for the regulatory effect of Bcl-xL-p53 interaction. According to Moll

et al., this interaction inhibits Bcl-xL and liberates Bak from Bcl-xL to engage MOMP. In

contrast, Chipuk et al. suggested that the cytoplasmic function of p53 is mediated by Puma

mediated loss of p53-Bcl-xL interaction (Chipuk et al., 2005).

The activity of p53 is mainly regulated by protein stability therefore it is not surprising

that the mitochondrial function of p53 is also shown to be regulated by stabilizing

modifications (Figure 2). Following stress, for example, an E3 ligase Mdm2-mediated

mono-ubiquitination has been suggested to promote mitochondrial translocation of p53

(Marchenko et al., 2007). Subsequently, p53 is immediately de-ubiquitinated by HAUSP

at the mitochondria, enabling interaction with Bcl-2 family members (Marchenko and

Moll, 2007). Moreover, Mdm2 antagonist Nutlin-3a liberates p53 by competitively

binding to the hydrophobic p53-binding pocket within the N-terminus of Mdm2 (Vassilev

et al., 2004). Furthermore, it has been suggested that Nutlin-3a could induce apoptosis via

a transcription-independent manner using mitochondrial p53 (Vaseva et al., 2009).

Controversial data has been published whether phosphorylation would be the determining

factor for mitochondrial translocation of p53 (Nemajerova et al., 2005; Park et al., 2005).

However, certain phosphorylation sites of p53 have been speculated to phosphorylated

and involved in the apoptotic function of p53 (Feng et al., 2006).

Page 45: Apoptosis by c-Myc - Helsinki

2.3.2 Death receptor apoptosis pathway On the surface of mammalian cells are receptors that are activated by death ligands such

as Fas/CD95, TNF-α, and TRAIL (Schmitz et al., 2000). These ligand bound death

receptors mediate apoptosis pathway by transmitting apoptotic signals to the caspases,

subsequently resulting in activation of caspase cascade and death (Ashkenazi and Dixit,

1998).

The major death receptors of the tumor necrosis factor receptor (TNFR) gene superfamily

include TNFR-1, Fas/CD95, DR (Death receptor)-4/TRAIL-R1, DR-5/TRAIL-R2 and less

known non-apoptotic DR3 and DR6 (Ashkenazi and Dixit, 1998; Pan et al., 1997). In

addition, osteoprotegerin is suggested to be one of the TRAIL receptors (Emery et al.,

1998). The death ligands of these receptors are members of the TNF family and play an

important role in the modulation of host defense mechanisms including: inflammation, T-

cell co-stimulation, induction of cell proliferation, macrophage activation, as well as

elimination of unwanted immune cells or tumor cells by apoptosis (Nagata, 1997).

All members of the TNFR family exist as mono- or trimeric receptors located on the

membrane and consist of cysteine-rich extracellular subdomains, which allow them to

recognize their ligands with specificity. Upon binding, trimerized death receptors expose

their conserved death domains (DD), which are located in the intracellular part of the

receptor (Chan et al., 2000). Different death receptors interact via these DDs with different

DD-containing adaptor proteins, like FADD, resulting in a DD-DD interaction and

formation of a death-inducing signaling complex (DISC) (Chinnaiyan et al., 1995). In

addition to its DD, the adaptor proteins also contain a N-terminal death effector domain

(DED), which allows further interaction with procaspase-8. CF95/Fas and TRAIL/Apo2L

ligands results in similar DISC formation and procaspase-8 processing and activation

(Figure 10) (Scaffidi et al., 1998). Signaling through a TNF receptor is more complex and

involves formation of two different complexes (I and II) and caspase-8 activation or

signaling via NF-κB or JNK (Micheau and Tschopp, 2003).

Page 46: Apoptosis by c-Myc - Helsinki

Figure 10. Death receptor signaling and apoptosis. When the CD95/Fas receptor is activated by its ligand CD95, it recruits FADD by DD-DD interaction resulting in DISC formation and DED exposure of FADD. Subsequently, FADD binds the pro-domain of caspase-8 via the DED and enables procaspase-8 dimerization, auto-activation and processing to p44/p41 and p12 subunits. TRAIL forms a DISC consisting of the activated receptor, FADD and caspase-8 or -10. However, the role of caspase-10 in induction of apoptosis is controversial. In TNF-a signaling the initial plasma membrane bound complex (complex I) consists of TNFR1, the adaptor TRADD, the kinase RIP1, and TRAF2 and its formation results in signaling via NF-kappa B or JNK. In a second step, TRADD and RIP1 associate with FADD and caspase-8, forming a cytoplasmic complex (complex II) leading to caspase-8 activation.

Page 47: Apoptosis by c-Myc - Helsinki

In addition, all DISC complex-mediated activations of caspase-8 can be inhibited by

DED-containing FLICE-inhibitory protein long form (FLIP), which consequently

interferes with FADD-caspase-8 interaction, inhibiting both caspase-8 activation within

the DISC and apoptosis (Micheau et al., 2002; Siegmund et al., 2001). Finally, the

activated caspase-8 can proceed to directly activate downstream caspase-3 or engage

mitochondrial apoptosis pathway.

In Type I cells, such as B-lymphocytes, activated caspase-8 can proceed to activate by

proteolytic processing the effectors caspase-3 and -7, which orchestrate apoptosis.

However, the majority of cells belong to Type II cells, like fibroblasts, where the ligand-

initiated apoptosis does not induce caspase activity strong enough for executing apoptosis,

thereby requiring amplification of the signals through the mitochondrial apoptosis

pathway (Rudner et al., 2005). The Bcl-2 family member Bid, which is a caspase-8 and -3

substrate, links the death receptor-induced caspase signaling to the mitochondria (Wang et

al., 1996). Activated caspase-8 cleaves Bid and its truncated form, tBID, translocates from

the cytosol to the mitochondria where it interacts with the pro-apoptotic Bcl-2 family

members Bax and Bak to induce the release of cytochrome c; whereby downstream

apoptotic events can occur (Figure 10) (Luo et al., 1998). Furthermore, the role of the

mitochondrial apoptosis pathway is evident in TRAIL-apoptosis, where inhibition of

TRAIL-induced apoptosis requires ablation of both Bak and Bax, as single knockouts are

still able to release cytochrome c (Kandasamy et al., 2003). However there exists

controversial data about whether Bax or Bak is the major effector in TRAIL-induced

apoptosis. In addition, Bcl-2 overexpression is able to inhibit TRAIL-induced apoptosis,

and interestingly, both Bid and caspase-8 cleavage were sown to be decreased due to Bcl-

2, indicating that caspase-8 activation is amplified by mitochondrial pathway (Fulda et al.,

2002). Also, Mcl-1 cleavage is required for sufficient TRAIL-induced cytochrome c

release (Weng et al., 2005). In addition, TRAIL may also induce activation of other

signaling pathways. For example, there exists evidence that JNK, p38 MAPK, and

IKK/NF-kB kinase pathways are activated downstream of DISC assembly and caspase-8

activation by TRAIL (Sheridan et al., 1997).

Page 48: Apoptosis by c-Myc - Helsinki

Interestingly, death receptor ligands were shown to exert toxicity specifically towards

cancer cells and, for example, recombinant TRAIL suppressed tumor growth without

apparent toxicity (Walczak et al., 1999). However, malignant tumor cells can escape from

treatments thus, a wide spectrum of studies has been focusing on defining the molecules

behind this resistance to treatment. Increasing clinical data however, indicates that TRAIL

is not effective against human cancers. Thus, already some studies have explored whether,

for example, DNA-damaging drugs could enhance the apoptotic effect of TRAIL

(Broaddus et el, 2005). Resistance to TRAIL-apoptosis in the cancer cell remains a

challenging issue for the successful application of TRAIL in cancer therapy.

The c-myc proto-oncogene encodes the c-Myc (herein Myc) transcription factor that was

originally identified as the cellular homologue to the viral oncogene (v-myc) of the avian

myelocytomatosis retrovirus (Vennstrom et al., 1982). c-Myc belongs to the family of Myc

genes (v-Myc, N-Myc, L-Myc, S-Myc and B-Myc) (Dang, 2012). In addition, Myc is a

64.5kDa transcription factor that contains a carboxy-terminal basic helix-loop-helix-zipper

(bHLHZ) domain and a transactivation domain (TAD) in its 150-amino acid amino-

terminus (Dang et al., 1989). In addition, the TAD contains highly conserved elements,

Myc Boxes (MB) I-IV, required for the transactivation of target genes. Myc activates a

diverse group of genes as part of a heterodimer complex with its partner protein, Max

(Blackwood and Eisenman, 1991). Myc-Max heterodimers are capable of binding specific

DNA sequences such as the E-box sequence (CACGTG) and subsequently activate

transcription of target genes (Adhikary and Eilers, 2005). Furthermore, Max association

and DNA binding are required for transcriptional activation of target genes by Myc as

well as its ability to drive proliferation, malignant cell transformation, and apoptosis

(Amati et al., 1993; Blackwell et al., 1990). Another protein family, the Mxd/Mad proteins

(Mad1, Mxi, Mad3, and Mad4) can antagonize Myc function by forming complexes with

Max. Recently, also other Mad-related proteins have been described (Mnt/Rox and Mga)

and some novel Max-related proteins (Mlx and Mondo) that interact with Mad-related

proteins (reviewed in Lutz et al., 2002). In addition, Myc has been suggested to regulate

Page 49: Apoptosis by c-Myc - Helsinki

chromatin by recruiting histone acetylation complexes via MBII-mediated binding to the

TRAP protein and via INI1, which interacts with the chromatin-remodeling complex

(Cheng et al., 1999; McMahon et al., 2000). Furthermore, Myc can act as a transcriptional

repressor with Max by binding directly MBII and the bHLHLZ domain, transcriptional

activators bound to DNA (Kleine-Kohlbrecher et al., 2006). Alternatively, to repress target

genes Myc associates with other proteins like Sp-1 and Miz-1 (Seoane et al., 2001)

Targeted gene disruption of both Myc alleles in embryonic stem cells leads to embryonic

lethality at day 9.5-10.5, which highlights the crucial role of Myc in normal growth

control during mammalian development (Davis et al., 1993). In approximately one-third

of human cancers, the expression of the Myc protein is deregulated, especially in

adenocarcinomas. Overexpression of Myc occurs through diverse mechanisms (discussed

at chapter 1.2.1) including translocation, amplifications, insertional mutagenesis, or

enhanced translation or protein stability (reviewed in Meyer and Penn, 2008). In addition,

point mutations in the coding sequence of the myc gene have been found in translocated

alleles of myc in Burkitt´s lymphoma. However, most often Myc expression is activated

indirectly through alterations in signaling pathways that induce or repress Myc

transcription or stability. Normally Myc transcription is regulated by elongation, and this

block is lost in cancer (Bentley and Groudine, 1986). Furthermore in non-transformed

cells the expression level of Myc is kept low by the extremely low half-life of its mRNA

and protein, approximately 30 minutes (Dani et al., 1984). More recently, it has become

clear that MYC can be deregulated by many additional mechanisms, including activation

by hormones or growth factors, their receptors, second messengers or transcriptional

effectors that enhance MYC expression. Additionally, in tumor cells the mRNA stability

of Myc is increased by direct and indirect mechanisms (Meyer and Penn, 2008). For

instance, increased Myc protein expression is caused by enhanced phosphorylation of

Ser62 by kinases that regulate the stability of Myc together with Thr58 which directs Myc

to proteosomal degradation (Hann, 2006). Furthermore, in cancer the stability of Myc is

enhanced by protein like Ras (Sears et al., 1999).

Myc binds to approximately 10-15% of the genome and is considered a global

transcriptional regulator (Figure 11). The discovery of Myc target genes started from using

fusion of Myc to the hormone-binding domain of the estrogen receptor (ER), MYC-ER,

Page 50: Apoptosis by c-Myc - Helsinki

and later on its tamoxifen regulated form (Eilers et al., 1989; Littlewood et al., 1995).

Importantly, large scale mRNA expression analysis of Myc target genes revealed that the

majority of Myc mRNA changes are relatively low, therefore deciphering true targets

requires more definition from assays such as the chromatin immunoprecipitation assay

(ChIP) (Meyer and Penn, 2008). Furthermore, more evidence suggests that Myc also

regulate target genes via inducing microRNAs (miRNA) from non-coding regions (Chang

et al., 2008).

3.2.1 Proliferation and differentiation by Myc Myc has a crucial role in cell proliferation that contributes to the neoplastic growth of

tumor cells. In quiescent cells in vitro, Myc expression is virtually undetectable. Either

after mitogenic stimulation or via ectopic expression of Myc, c-Myc mRNA and protein

are rapidly induced and subsequently, the cell enters the G1 phase of the cell cycle (Eilers

et al., 1989; Rabbitts et al., 1985). In the cell cycle, Myc is shown to be essential for

GO/G1 to S phase progression and G1 seems to be shortened in the Myc-expressing cells.

Furthermore, Myc allelic deficiency results in defects in cell growth, such as cell mass,

and in the G1 phase (reviewed in Obaya et al., 1999). The cell cycle is regulated by

multiple target genes of Myc. For example, Myc abrogates the transcription of cell cycle

checkpoint genes (GADD45, GADD153) and inhibits the function of cyclin-dependent

kinase (CDK) inhibitors. In addition, Myc promotes cell cycle progression by activation of

cyclin D1, -D2, -E1, -A2, CDK4, Cdc25, E2F1, and E2F2 (Meyer and Penn, 2008).

Recent studies also suggest that Myc might influence the cell cycle by repressing genes,

such as the cell cycle arrest-inducing genes p15 and p16 (Staller et al., 2001). Importantly,

Myc allows cells to enter the S phase and undergo mitosis in the absence of external

growth factor (Eilers et al., 1991).

Myc is broadly expressed during embryogenesis and its expression correlates with growth

in the growing tissues of adults (Pfeifer-Ohlsson et al., 1985). Furthermore, differentiation

is regulated by Myc in both ways; Myc can block the differentiation of a variety of cell

types. In contrast, MYC has been suggested to stimulate cellular differentiation in the case

of stem cells (Gandarillas and Watt, 1997; Langdon et al., 1986). Importantly, the role of

Page 51: Apoptosis by c-Myc - Helsinki

Myc-induced de-differentiation seems to have a major effect on tumor cell phenotype.

Recently, Myc has been shown to increase the capacity of stem cell self-renewal and to

induce pluripotent stem cells (iPS) in terminally differentiated cells (Okita et al., 2007;

Wilson et al., 2004).

Figure 11. Myc-induced molecular and cellular changes.

3.2.2 Transformation mechanisms by Myc The mechanism behind the tumorigenic function of Myc considers proliferation and

differentiation, but also the ability of Myc to increase cell growth (cell mass in size). This

occurs by increasing protein synthesis and reprogramming cell metabolism (Iritani and

Eisenman, 1999). In accordance, enforced expression of the Drosophila c-Myc ortholog in

wing imaginal disc results in increased cell size (Johnston et al., 1999). Importantly,

genomic instability and amplification are increased in Myc-expressing cells, for example

by ROS generation or overcoming the p53 checkpoint (Felsher and Bishop, 1999). More

recently, Myc was suggested to regulate angiogenesis by inducing angiogenic switch via

miR17-92 microRNA cluster or interleukin 1β (Dews et al., 2010; Shchors et al., 2006).

Page 52: Apoptosis by c-Myc - Helsinki

Additionally, Myc can induce immortalization of a cell, however, the mechanisms are

very controversial (Lutz et al., 2002). Specific metabolic pathways are also targeted by

Myc, as discussed in chapter 3.4. Furthermore, Myc is downstream from many signal

transduction pathways like MEK-ERK and PI3K that are responsive to growth factors or

the cellular microenvironment (Lee et al., 2008). Conversely, Myc tumors have been

shown to be resistant to anti-mitogenic TGF-β signaling (Meyer and Penn, 2008).

Several transgenic mouse models have been developed to decipher the mechanism how

deregulated Myc induce tumorigenesis. The mouse data has provided the evidence that

deregulated expression of Myc is sufficient to drive tumorigenesis in a number of

transgenic mouse tissues, but not all (Adams et al., 1985; Leder et al., 1986). For instance,

Myc induced tumor in mammary mouse tissue under Wap- or MMTV-promoter and in

mouse prostate tissue, however, with delayed onset of tumors (Ellwood-Yen et al., 2003;

Schoenenberger et al., 1988; Stewart et al., 1984). Importantly, in each of these cases

additional mutagenic events are necessary for Myc to enable its tumorigenic potential.

Importantly, using inducible systems (tTA Tet-O-Myc, pIns-MycERtm) it has become

evident that withdrawal of Myc ectopic expression is required to maintain tumor, or

otherwise the tumor will regress (Arvanitis and Felsher, 2006; Karlsson et al., 2003).

Importantly, these tumors are considered to be addicted to Myc. Furthermore, increasing

data on mouse models suggest that evading Myc-apoptosis is one of the major

mechanisms behind the oncogenic function of Myc (Dang et al., 2005).

Several laboratories made an intriguing discovery in the early 90´s that oncogenes such as

Myc and the adenovirus E1A, which are both potent inducers of cell proliferation, induce

also apoptosis. The early view of oncogene-induced apoptosis was that it was indirect or

distal response of the cell to an enforced cell cycle entry or to an inappropriate growth

signal generated by Myc. Currently it is evident that Myc-induced apoptosis has a close

link to the Bcl-2-mediated mitochondrial apoptosis pathway. Originally Bcl-2 was found

to cooperate with Myc in tumorigenesis by inhibiting Myc-induced apoptosis (Fanidi et al.,

1992). Similarly, acceleration of lymphomagenesis is found in transgenic mice that

express both Myc and Bcl-2 compared with transgenic mice that harbor just the Myc

Page 53: Apoptosis by c-Myc - Helsinki

transgene (Strasser et al., 1990). Also Myc-null cells were resistant to diverse apoptotic

stimuli. Furthermore, ectopic expression of Myc in the absence of specific survival factors

lead to apoptosis (Askew et al., 1991; Evan et al., 1992). Later Myc was shown to induce

the release of cytochrome c from mitochondria during apoptosis (Juin et al., 1999).

Several studies suggest an important role for Bax in Myc-mediated engagement of the

mitochondrial pathway. For instance, Myc was shown to be required for DNA damage-

induced Bax conformational change and activation and also to Bax oligomerization at

membranes. (Annis et al., 2005; Soucie et al., 2001). Interestingly, in the pIns-MycERtm

pancreas model, the Myc overexpression is able to induce apoptosis of β-cells

simultaneously with the proliferation. However, expression of Bcl-xL or loss-of Bax are

shown to enable Myc-tumorigenesis in these β-cells (Dansen et al., 2006; Pelengaris et al.,

2002). Also, in the MMTV-Myc mammary transgenic mouse model, Myc induces

apoptosis in a Bax-dependent manner. Additionally, Myc has been shown to up-regulate

Bax in growth factor-deprived human cancer cell lines (Mitchell et al., 2000). In non-

epithelial cell tumor models, Myc can co-operate with Bcl-2 and Bcl-xL overexpression in

Εμ-Myc lymphoma and with loss of Bim in Εμ-Myc B-cell leukemia (Egle et al., 2004).

Some studies have shown evidence that Myc suppresses Bcl-2 or Bcl-xL expression both

in MEFs and primary hematopoietic cells (Eischen et al., 2001b; Kelly et al., 2011).

Furthermore, Myc has been proposed to affect the Bcl-2 family network controlling the

mitochondrial pathway, yet the mechanisms by which Myc actually increase the activation

of Bax and Bak remain unknown.

Importantly, tumor suppressor p53 seems to have an essential for Myc-induced apoptosis

(Hermeking and Eick, 1994). For instance, elevated expression of p53 appears to be

associated with apoptosis occurring by Myc (Hermeking and Eick, 1994). Myc-induced

phosphorylation of Serine 15 (murine homologue 18) inhibits Myc tumorigenesis in Εμ-

Myc mice (Sluss et al., 2010). Furthermore, p53 phosphorylated by ATM promotes Myc

apoptosis in squamous epithelial cells (Pusapati et al., 2006). Likewise, Myc co-operates

in tumorigenesis with loss of p53 or ARF in variety of models like in lymphomas, but not,

for example, in mammary carcinomas (Eischen et al., 1999; Elson et al., 1995).

Conversely, existing data in several systems suggest that elevated Myc can induce

apoptosis in the absence of p53, for example, by suppressing Bcl-xL. Thus, Myc-induced

apoptosis is not always dependent on p53 (Hsu et al., 1995; Maclean et al., 2003).

Page 54: Apoptosis by c-Myc - Helsinki

Similarly, Myc can also induce p53 through p14ARF-independent mechanisms in human

fibroblasts (Lindstrom and Wiman, 2003). Importantly, the Myc binding partner Max

seems to be necessary for apoptosis following inappropriate Myc expression, indicating a

transcriptional role in apoptosis (Amati et al., 1993).

Furthermore, increased CDK2 kinase activity is found from Myc-expressing apoptotic

cells and it has been shown to be required for Myc-induced apoptosis in MEFs. Cyclin D3

also sensitizes cells to TNF-induced Myc-dependent apoptosis (Janicke et al., 1996). In

contrast, the mRNA of cyclin A is up-regulated in cells overexpressing Myc but the CDK

activity and apoptosis were shown to be independent events that occur in response to

activation of Myc (Rudolph et al., 1996). Yet the exact role of CDKs in Myc-induced

apoptosis is still unknown. Additionally, the apoptotic properties of Myc are blocked by

cytokines and adhesion signals. For instance, survival signals, either from Ras signaling of

IGF1 receptor, have been shown to inhibit Myc-induced apoptosis e.g. by blocking

cytochrome c release (Kauffmann-Zeh et al., 1997). Whereas, for example, TGF-α

appears to suppress apoptosis by Myc during development of mammary adenocarcinoma

(Harrington et al., 1994). In addition, cell adhesion also influences the susceptibility of

cells to apoptosis by Myc (McGill et al., 1997). Further studies are needed to assess

whether loss of integrin-mediated adhesion is a triggering event or a correlate of apoptosis

by Myc.

Expression of Myc sensitizes cells to a wide range of other pro-apoptotic stimuli, such as

hypoxia, DNA damage, and depleted survival factors as discussed earlier, as well as to

signaling through the Fas/CD95, TNF-α, and TRAIL death receptors (Hueber et al., 1997;

Klefstrom et al., 1994; Ricci et al., 2004b). The sensitization to death receptors signaling

is suggested to occur immediately downstream of these receptor complexes, for example,

by involving p53 and blocking activation of NF-κB or by targeting RIP in TNF-α

signaling (Klefstrom et al., 1997; Klefstrom et al., 2002). Furthermore, one cancer cell line

study suggests that Myc suppressed FLIP and thereby increases caspase-8 processing at

the DISC (Ricci et al., 2004b). Altogether, there seem to be a wide variety of different

ways for Myc to engage apoptosis and this can be explained, for example, by tissue

specificity.

Page 55: Apoptosis by c-Myc - Helsinki

After finding ways to enable transcription and growth, Myc has to organize energy for the

multiplying cell. Glucose and glutamine are the main nutrients that cultured mammalian

cells catabolize for energy production and biosynthesis (Vander Heiden et al., 2009;

Yuneva et al., 2007). Myc directly induces up-regulation of proteins responsible for

glucose uptake and glycolysis (Figure 12). Moreover, Myc has also been implicated in

promoting RNA splicing for the expression of pyruvate kinase M2 that slows down

pyruvate formation and direct glucolytic intermediates to be used in biosynthetic pathways

(David et al., 2010). Moreover, Myc stimulates glutamine uptake and metabolism by

elevating expression levels of glutamine transporters and glutaminase (GLS) via miR23a

and miR23b, (Chang et al., 2008; Gao et al., 2009; Wise et al., 2008). Thereby,

deregulated Myc reprograms metabolism toward using glutamine as the main oxidizable

substrate for maintenance of the tricarboxylic acid cycle and ATP production, thereby

stimulating glutamine consumption and metabolism and also generation of reactive

oxygen species (Gao et al., 2009; Wise et al., 2008). In addition to providing carbon for

bioenergetic reactions, glutamine is also the obligate donor of nitrogen in the biosynthesis

of nucleic acids, and the primary source of nitrogen for the synthesis of non-essential

amino acids (Wise and Thompson, 2010). More importantly, Myc-overexpressing cells

depend on a continual supply of nutrients thereby causing them to be addicted, for

example, to glutamine (Wise et al., 2008). Furthermore, glutamine metabolism appears to

be important for cell survival under especially glucose- or oxygen-deprived conditions (Le

et al, 2012). Another feature of cancer cells is inefficient way of producing ATP, since

nutrients are being incorporated into biomass. In accordance, Myc-mediated metabolic

reprogramming has been associated with lower ATP levels (Vander Heiden et al., 2009;

Yuneva et al., 2007). Concomitantly, Myc is shown to activate genes involved in glucose

and glutamine usage for nucleotide biosynthesis (Mannava et al., 2008). For instance, Myc

induces glycolytic flux from 3-phosphoglycerate for the synthesis of serine and glycine,

which are essential for nucleotide biosynthesis (Vazquez et al., 2011). Glucose and

glutamine are also essential for fatty acid synthesis and Myc is suggested to incorporate

glucose carbons into fatty acids (Morrish et al., 2010). Altogether, these Myc-induced

molecular changes in glycolysis and glutaminolysis are reflecting cell´s adaptation to

abnormal proliferation and requirement of new building blocks.

Page 56: Apoptosis by c-Myc - Helsinki

Figure 12. Metabolic reprogramming by Myc and the metabolically tumor suppressive function of p53 Myc increases glycolysis by upregulating the expression of glucose transporters proteins Glut1 and Glut4, hexokinase (HK2) and pyruvate dehydrogenase kinase 1 (PDK1). Myc increases glutaminolysis by enhancing the expression of glutamine transporters SCT2, SN2 and SLC7A5 and GLS enzyme. Serine and Nucleotide biosynthesis is enhanced by Myc through increased glycolytic flux and upegulation of proteins like PKM2. Mitochondrial biogenesis is affected by Myc via PGC-1β. P53 counterbalances metabolic effects of Myc via increasing OXPHOS, upregulating GLS and inhibiting glycolysis by TIGAR. Modified from (Li and Simon, 2013).

Importantly, Myc regulates mitochondrial biogenesis by increasing oxygen consumption

and mitochondrial mass and function (Li et al., 2005). In accordance, Myc can directly

reactivate genes involved in mitochondrial biogenesis and in mitochondrial protein

synthesis (Ahuja et al., 2010). Additionally, Myc has been suggested to regulate

mitochondrial fusion and fission (Graves et al., 2012). Intriguingly, ARK5, an AMPK

family member, is shown to be essential for Myc-induced metabolic reprogramming and

Page 57: Apoptosis by c-Myc - Helsinki

energy homeostasis, since ARK5 deficiency is lethal with Myc expression (Liu et al.,

2012). All together, recent data emphasizes how widely Myc is affecting energy

homeostasis, however, future studies are needed to fully understand how metabolic

alteration are linked to activation of apoptotic pathways by Myc.

Page 58: Apoptosis by c-Myc - Helsinki

AIMS OF THE STUDY

The specific aims were:

I To elucidate the role of mitochondrial apoptosis pathway engagement in Myc-induced

apoptosis and analyze the potential re-activation of apoptosis in mammary Myc-tumors.

II To distinguish and characterize the crucial mechanisms of Myc-induced Bak activation

and conformational change

III To explore connections of the Myc-mediated mitochondrial apoptosis pathway and

metabolic reprogramming.

Page 59: Apoptosis by c-Myc - Helsinki

MATERIALS AND METHODS

The material and methods used in this study are listed below and are described in detail in

the original publications, which are here referred to using Roman numerals.

1. Materials

Mouse lines Description Source of reference used in

FVB Inbred mouse strain Harlan/Meilahti III, IV

animal facility

Wap-Myc (FVB.Cg-Tg Mouse ine expressing c-Myc under Jax Laboratories III, IV

(Wap-Myc)212Bri/J Whey acidic protein (wap) promoter (stock #002677)

Cells HEK-293-fT Human embryonic kidney fibroblast ATCC I-III

immortalized by hTERT

BT-474 human breast cancer cell line ATCC III

CCD 1102 KERTr keratinocyte ATCC III

HeLa cervical cancer ATCC III

HMEC-hT primary human mammary epithelial ATCC I-III

cell immortalized by hTERT

HepG2 hepatocellular carcinoma ATCC III

MCF10A non-transformed human mammary ATCC I-IV

epithelial cell line

MCF12A human mammary epithelial cell line ATCC III

MCF7 human breast cancer cell line ATCC III

MDA-MB-361 human breast cancer cell line ATCC III

MDA-MB-436 human breast cancer cell line ATCC III

MRC5-hT primary human lung fibroblasts ATCC I, III

immortalized by hTERT

RPE retinal pigmented epithelial cell ATCC III

Sk-Br3 human breast cancer cell line ATCC III

Phoenix Ampho retroviral packaging cell line Dr. Garry Nolan I, II

NMuNG mouse mammary epithelial cell line ATCC IV

ZR-75-30 human breast cancer cell line ATCC III

Page 60: Apoptosis by c-Myc - Helsinki

Expression vectors Description Source of reference used in

pcDNA3.1-caspase-8 encodes caspase-8 provided by Dr.Evan I

pcDNA3.1-caspase-8 C-A encodes catalytic site C->A mutant provided by Dr.Evan I

caspase-8

pC4M-Fv2E- caspase-8 protease domain of casp-8 cloned Prepared in Klefström I

into a FKBP variant construct allowing laboratory

dimerization

pC4M-Fv2E- caspase-8 protease domain of catalytic site Prepared in Klefström I

C->A mutant laboratory

pcDNA3.1-Mycwt encodes Myc provided by Dr.Evan I

Viral vectors Description Source of reference used in

pBabe-puro-MycERtm retoriviral expression vector encoding Myc provided by Dr.Evan I-III

fused with hormone binding domain 4-OHT

pBabe-puroMydcER retoriviral expression vector encoding Myc provided by Dr.Evan I-III

with deletion in aminoacids 106-143

fused with hormone binding domain 4-OHT

pBabe-hygro-Bcl-xL retroviral expression vector encodes provided by Dr.Evan I, II

Bcl-xL

pWZL-hTERT retroviral expression vector endocing provided by Dr.McMahon I-III

catalytic subunit of human telomerase I

pENTR-H1 Gateway entry vector Biocentrum Helsinki I-III

Genome Biology unit

pDSL-hpUGIH Lentiviral shRNA vector with GFP marker Alliance for Cellular I-III

and ires hygromycin selection marker Signalling (AfCs)

pLentiLox3.7 (pLL3.7) Lentiviral shRNA vector with GFP marker Dr. Luk Van Parijs I, II

pLenti6/V5-p53WT Lentiviral expression vector with V5 tag Addgene III

encoding p53 wt

pLenti6/V5-p53R175H Lentiviral expression vector with V5 tag Addgene III

encoding transcription deficient mutant

Addgene III

Page 61: Apoptosis by c-Myc - Helsinki

shRNA vectors antisense sequence (5´-3´´) Source of reference used in

pLL3.7 shcontrol TTAGTAGAGATCTATGAGA prepared by author I, II

pLL3.7 shcaspse-8 CAGGGTCATGCTCTATCAGATTT prepared by author I, II

pLL3.7 shLkb1-1 CCCAAGGCTGTTTGTGTGAAT prepared in Klefström lab II

pLL3.7 shLkb1-2s GCAGAAGATGTATATGGTGAT prepared in Klefström lab II

pDSL-UGIH-shbak CATAGCGTCGGTTGATGTCgt prepared by author I, III

pDSL-UGIH-shbid ACAGAATCCGCACTGTGCT prepared by author I, II

pDSL-UGIH-shbim TGTTGATTTGTCACAACTC prepared by author II

pDSL-UGIH-shbcl-xL UUCACUACCUGUUCAAAGCTT prepared by author III

pDSL-UGIH-shmcl-1 UUCGAUGCAGCUUUCUUGGTT prepared by author III

pDSL-UGIH-sha1 UAUCCAAAUUCACAGUCUGTT prepared by author III

Antibody Description Source of reference used in

α-actin mouse monoclonal Sigma I, II

α6-integrin mouse monoclonal Cymbus Biotechnnologies II

β-actin Santa Cruz III

ACC rabbit polyclonal Cell Signaling Technology III

p-ACCSer79 rabbit polyclonal Cell Signaling Technology III

AMPK rabbit polyclonal Cell Signaling Technology III

p-AMPK Thr172 rabbit polyclonal Cell Signaling Technology III Bak (N-terminal, ab-1) Calbiochem I, III

Bax (N-terminal) rabbit polyclonal Cell Signaling Technology 1, III

Bax (N-terminal, 6A7) mouse monoclonal Trevigen I

Bcl-xL mouse monoclonal BD Transduction Lab. I-III

Bcl-xL rabbit polyclonal Cell Signaling IV

Bim rabbit polyclonal Cell Signaling I, II

Bid rabbit polyclonal Cell Signaling I,

β-catenin mouse monoclonal BD Transduction Lab. II

caspase-3 (Asp175) rabbit polyclonal Cell Signaling Technology I-IV

Caspase-8 IC12 mouse monoclonal Cell Signaling Technology I

CK14 rabbit polyclonal Covance IV

c-Myc (9E10) mouse monoclonal Covance I, III

Cytochrome c mouse monoclonal Pharmingen I, II

COXIV mouse monoclonal Molecular Probes III

Collagen IV rabbit polyclonal Abcam IV

E-cadherin mouse monoclonal BD Transduction Lab. II

EpCAM rabbit polyclonal Abcam IV

GAPDH mouse monoclonal Cell Signaling Technology III

Glutaminase rabbit polyclonal Abcam III

Page 62: Apoptosis by c-Myc - Helsinki

Antibody Description Source of reference used in

GM-130 mouse monoclonal Cell Signaling Technology II

Histone 3 mouse monoclonal Santa Cruz III

Hsp70 mouse monoclonal Santa Cruz III

Ki67 (NCL-Ki67p) rabbit polyclonal Novocostra Lab III

Ki67 (M724901 rat monoclonal DAKO IV

Lamin B goat polyclonal Santa Cruz III

Lkb1 mouse monoclonal Abcam II

Mcl-1 S-19 rabbit polyclonal Santa Cruz III

Mcl-1 600-401 rabbit polyclonal Rockland III

P21 Cell Signaling Technology III

PCNA rabbit polyclonal Santa Cruz III

Phospho S6 Ser223/236 rabbit polyclonal Cell Signaling Technology II

Puma (ab9643) rabbit polyclonal Cell Signaling Technology I

P53 Ab-1 Calbiochem III

P53 Bp53-12 rabbit polyclonal Santa Cruz III

P53 FL-393 mouse monoclonal Santa Cruz III

p-p53 Ser15(18) rabbit polyclonal Cell Signaling Technology III

β-tubulin, rabbit polyclonal Abcam III, IV

glutaminase rabbit polyclonal Abcam III

Page 63: Apoptosis by c-Myc - Helsinki

2. Methods

The following section describes in detail the main methods used by the author.

Administration of ABT-737 to WAP-Myc mice and measure of tumor growth (IV)

WAP-MYC females mice underwent two pregnancies after 8 week of age. ABT-737 was

first diluted in 65% D5W carrier solution (pH 3-4). Animals received either ABT-737

100mg/kg/d solution or equal volume of pure carrier solution administrated by

intraperitoneal injections (i.p) for 21 consecutive days. After treatment period mice were

followed for another 21 days. To measure the tumors, experimental animals were

monitored twice a week in three-day intervals with an electronic caliper (Clas Ohlson).

The formula used for calculating tumor volume is V= d2* D/2, where d= shortest diameter

and D= longest diameter

Cell culture (I‐IV)

Mammary epithelial MCF10A and HMEC-hT cells were cultured in epithelial cell basal

media MCDB 170 (US Biological) supplemented with insulin 5 μg/mL, Bovine Pituitary

Extract (BPE) 70 μg/mL, Hydrocortison 0.5 μL/ mL, Epidermal Growth Factor 5 ng/mL,

Human Transferrin 5 μg/ mL, and Isopropretenol 0.01 μM. MCF12A cells (ATCC) were

cultured in DMEM/F-12 (Gibco) supplemented with 10% (vol/ vol) FCS, epidermal

growth factor 20 ng/mL, hydrocortison 1 μg/mL, insulin 10 μg/mL, human transferrin 5

μg/mL, glutamine, and antibiotics. Mammary breast adenocarcinoma Sk-Br3 cells were

grown in McCoy_s 5a media (Lonca), MDA-MB-361 and MDAMB- 436 in Leibovitz_s

media (Sigma), BT-474, HeLa, HepG2, and MRC-5-hT cells in DMEM, ZR-75–30 cells

in RPMImedium1640 (Lonza), RPE cells in DMEM/F-12 (Gibco), and MCF7 in EMEM

(Lonza) all supplemented with 10% FCS, glutamine, and antibiotics. MDA-MB-436 and

MCF-7 media had 10 μg/mL insulin. CCD 1102 KERTr (Keratinocyte) cells where

cultured with Keratinocyte- SFM Medium (Kit) with L-Glutamine, EGF, and BPE

(Gibco). Phoenix Ampho packaging cells were cultured in high‐glucose containing D‐MEM (Gibco) supplemented with 10% FCS, glutamine and antibiotics (penicillin and

streptomycin). Primary WAP-MYC mouse mammary tumor cells were cultured in .

In three-dimentional organotypic cell culture, the cells were grown in Matrigeltm, which is

Page 64: Apoptosis by c-Myc - Helsinki

a Basement membrane from Engelbreth‐Holm‐Swarm mouse sarcoma (Becton

Dickinson). The mat was prepared according to manufacturers instructions and Matrigel

was overlaid on 8‐chamber slides and left to solidify for 20 minutes at +37°C.

Subsequently MCF10A cells were trypsinized from 2D culture and resuspended in assay

media (MCDB 170 growth media with supplements containing 2% Matrigel) for seeding

(1500 cells/well) on top of the Matrigel. To prepare Collagen I gel 3D cultures, 3 mg/ml

of Collagen I gel (Rat tail Collagen I, Cultrex) and 23 mM NaOH were mixed on ice with

10x phosphate buffered saline (PBS) and subsequently, the mixture was overlaid on 8‐chamber slides and let to solidify in +37°C for 1 hour. Cells (1500 cells/well) were plated

onto Collagen in MCDB media with supplements.

Cloning of shRNA (I, II)

To silence caspase-8 60 bp shRNA oligos were designed and ordered to target caspase-8

mRNA and one scramble (non-targeting control) (Qiagen). The oligonucleotides were

annealed in Annealing Buffer as follows: at +95 °C for 4 min, +70 °C for 10 min and then

temperature was decreased to + 4 °C slowly (6 °C/min). Annealed oligos were clones into

lentiviral vector LentiLox 3.7 (pLL3.7). To silence Bcl-2 family members shRNA oligo

sequences of mRNA designed by Biomedicum Bioinformatics Unit and Biocentrum

Helsinki Systems Biology Initiative and ordered from (Proligo). The annealed shRNA

oligos were cloned first into pENTRY Gateway vector and subsequently using gateway

cloning into lentiviral pDSL_UGIH.

Chemical Cross-Linking (III)

Cells were collected from 15-cm plates into ice-cold PBS and pelleted by centrifugation at

1,500 rpm (Heraeus Megafuge 1.0 R) for 10 min. The cell pellet was resuspended in

SCEB buffer (200 mM sucrose, 10 mM Hepes pH7.4, 50 mM KCl, 5 mM EGTA, 5 mM

MgCl2,) containing DTT, PMSF, and cytochalasin B and incubated for 30 min before

homogenization with Dounce homogenizer. The unbroken cells and debris were removed

by centrifugation at 2,000 rpm (Heraeus Megafuge 1.0 R) for 10 min at +4 °C.

Subsequently, the supernatant was subjected to a cross-linking assay. BMH (1,6-

bismaleimidohexane) was added to a final concentration of 1 mM, and lysates were

incubated for 30 min at room temperature on constant mixing. After quenching the

reaction with DTT, the mitochondria-enriched heavy membrane fraction was pelleted at

Page 65: Apoptosis by c-Myc - Helsinki

13,000 rpm (Heraeus Pico17) for 10 min at +4 °C and lysed in radioimmunoprecipitation

assay buffer (RIPA). From each sample, 20 μg of mitochondrial proteins was subjected to

SDS/PAGE and Western blot analysis.

Co-immunoprecipitation (III)

Cells growing on 15-cm dishes were lysed in 300 μL of IP-lysis buffer (20 mM Tris·HCL

pH 7.4, 135 mM NaCl, 1.5 mM MgCl2, 1 mM EGTA, 10% (vol/vol) Glycerol, 1%

(wt/vol) CHAPS with protein inhibitor mixture (Roche). Lysates were kept on ice for 10

min, and unbroken cells and debris were centrifuged at 13,000 rpm (HeraeusMegafuge 1.0

R) for 15 min at +4 °C. Immunoprecipitations were carried out using the Catch and

Release v2.0 Reversible Immunoprecipitation System kit (Millipore). Columns were

washed according to the protocol, and the reaction mix for immunoprecipitation contained

2 μg of IP-antibody, 2,000 μg of protein, and antibody capture affinity ligand. Samples

were incubated on rotation at +4 °C for 2 h and washed five times with 400 μL of Catch

and Release Washing Buffer. Antibody captured proteins were eluted with 45 μL of Catch

and Release Denaturing Buffer containing 200 μL of β-mercaptoethanol and by

centrifuging at 5,000 rpm (Heraeus Pico17) for 20 s.

Caspase-8 and caspase-3 activity assay (I, III, IV)

To measure caspase activities (I), the tells were lysed in Triton-X cell lysis buffer (10mM

Tris, pH 7.5 130mM NaCl, 1% Triton X-100, 10mM NaPi and 10mM NaPPi) and the

amount of caspase-8 activity was estimated by measuring proteolytic activity towards

synthetic caspase-8 substrate Ac-IETDpNA or caspase-3, Ac-DEVD-pNa peptide

(Calbiochem). To measure caspase-8 or caspase-3 activity, 40 mg aliquot of each lysate

was incubated in a 100 ml total volume of 1_Salvesen optimal caspase buffer (Stennicke

and Salvesen, 1997) supplemented with 200 mM colorimetric pNa-peptide. Samples were

incubated at 37 °C and read at 405nm with an automated plate-reading spectrophotometer

at 1-h intervals. In addition 96-well commercial kit was used (III, IV) and ells were seeded

on 96-well plates and treated, and apoptosis was measured using the Caspase-Glo 3/7

luminescence kit. Luminescence (RLU) was measured using a VICTORX3 (PerkinElmer)

plate reader.

Page 66: Apoptosis by c-Myc - Helsinki

Generation of Fv-caspase-8 constructs (I)

The protease domains of caspase-8 and caspase-8 C-A mutant (catalytic site cysteine

mutated into an alanine) were PCR amplified from pcDNA3.1-Caspase 8 and pcDNA3.1-

Caspase 8 C-A vectors. Resulting fragments were digested and subcloned into pC4M-

Fv2E (Ariad Pharmaceuticals). The two Fv (FKBP variant) domains in the chimeric Fv2E-

caspase-8 construct substitute for the natural DED dimerization domains of caspase-8 and

allow ectopic dimerization of the construct with chemical dimerizer AP20187. The

sequences of pC4M-Fv2E-caspase-8 and pC4M-Fv2E-caspase-8 C-A constructs were

verified by sequencing.

Immunofluorescence of two-dimensional cultures (I‐III) All cells were fixed with 4% paraformaldehyde (PFA) for 10 minutes at room temperature

(RT) before staining. To detect localization of cyt c, fixed cells were washed for 6 times

with PBS to permeabilize and block with PBS containing 3% BSA and 0.05% Saponin.

Next, the cells were incubated for 1h at RT with monoclonal anti-cyt c antibodies (BD

Pharmingen) diluted 1:300 in permeabilization/blocking buffer. After incubation with

primary antibody, cells were washed for 6 times and incubated for 1h with appropriate

Alexa fluor dye (Alexa 488 or 546, Molecular Probes) coupled secondary antibodies. Next,

cells were washed with PBS and counterstained with 1 μg/ml Hoechst 33258 (Sigma) to

visualize the nuclei. For immunostainings with conformation-specific antibodies (Bak and

Bax), cells were were blocked and permeabilized with 5% goat serum+0,06% Saponin and,

the primary and secondary antibodies were diluted 1:300 in 5% goat serum+0,04 %

Saponin. All other staining were permeabilized with 0.1% (vol/vol) Triton X-100 BSA-

PBS for 30 min and primary and secondary antibodies were incubated in 1% (wt/vol)

BSA-PBS for 1 h at room temperature and washed with PBS. Hoechst 33258 (Sigma) was

used to detect the nuclei. Slides were mounted using Immu Mount (Thermo Scientific)

mounting reagent. Images were acquired using Zeiss Axioplan 2 microscope equipped

with mercury lamp and Zeiss AxioCam HRc color camera and Axiovision 4.6 software

using objectives Plan Neofluar 20x (NA=0.5) and 40x (NA=0.75), 64x. (NA=1.25).

Immunofluorescence of three-dimensional cultures (II)

The organotypic cultures were fixed with 2% PFA for 20 minutes at RT and thereafter

washed with PBS. Epithelial structures were permeabilized with 0.25% Triton X‐100 in

Page 67: Apoptosis by c-Myc - Helsinki

PBS for 10 minutes at +4°C, and thereafter washed with PBS. The non‐specific binding

sites were blocked in immunofluoresence (IF) buffer (7.7 mM NaN3, 0.1% BSA, 0.2%

Triton X‐100 (Sigma) and 0.05% Tween20 in PBS) supplemented with 10% normal goat

serum (Gibco) for 1‐2 hours. Subsequently, slides were incubated with primary antibody

in the blocking solution overnight at +4° following three 15 minutes washes with IF buffer.

Then slides incubated with Alexa Fluor secondary antibody diluted in blocking solution

for 40‐50 minutes at RT following washes with IF buffer. The nuclei were counterstained

with Hoechst33258. Slides containing immunostained epithelial structures were mounted

with Immu‐Mount mounting reagent. Images of the structures were acquired using Zeiss

Axiovert 200 microscope equipped with Zeiss Apotome and AxioCam MRm camera

(study IV) and/or Zeiss LSM Meta 510 confocal microscope (studies I‐IV) equipped with

argon (488), helium‐neon (543 and 633) and diode (405) lasers. The objectives used were

Plan‐Neofluar 40x DIC objective (NA=1.3, oil) and Plan‐Neofluar 63x DIC (NA=1.25,

oil).

Immunohistochemistry (III, IV)

Mammary gland tissue and tumors were collected and fixed in 4% (vol/vol)

paraformaldehyde (PFA) overnight and processed into paraffin blocks. Sections (3 μm)

were cut from paraffin blocks on glass slides (Superfrost Plus), deparaffinised and

incubated overnight at +70 °C in antigen retrieval citrate buffer solution (Dako).

Subsequently, sections were incubated in 0.3% (vol/vol) H2O2 solution for 15 min and

washed for 5 min with PBS. Sections were blocked in buffer containing 10% (vol/vol)

normal goat serum (Gibco) in mammary gland histological buffer [0.1% (wt/vol) BSA,

0.2% (vol/vol) Triton-X (Sigma), 0.05% (vol/vol) Tween 20] for 30 min followed by 1 h

incubation with primary antibody and secondary antibody using standard avidin–biotin

complex detection with 3,3-diaminobenzidine (Vector Laboratories), and counter- staining

with hematoxyline (Thermo Scientific). Sections were dehydrated and mounted with

Vectastain permanent mounting medium (Vector Laboratories). TUNEL staining was

performed with ApopTag in situ detection kit (Chemicon/Millipore) according to

manufacturers instructions. For all paraffin embedded tissues, hematoxylin and eosin

(H&E) staining were performed in Biomedicum Genomics Histolab. The

immunohistochemistry (IHC) samples were analyzed and photographed with a Leica DM

Page 68: Apoptosis by c-Myc - Helsinki

LB microscope and software Studio-Lite 1.0.

Primary cell isolations and fat pad transplantations (IV)

To isolate primary WAP-Myc mouse mammary epithelial cells were isolated from 8‐14

weeks old virgin female mice. To isolate tumor cells, mice were euthanized with CO2 and

cervical dislocation and thoracic and tumors were removed and chopped.

(Heraeus Multifuge 1.0)

0.05% Trypsin‐EDTA for 5‐10 minutes

Fat pad transplantations were performed essentially as described previously (Deome et al.,

1959; Welm et al., 2008). 3‐week‐old female FVB mice were anesthetized using 2‐2.5%

Isoflurane (Baxter) and the anterior part of #4 mammary gland containing rudimentary

ductal epithelium, lymph node and bridge to #5 gland was surgically removed. Wap-Myc

tumor 105 cells/gland in 10 μl volume were injected to the remaining fat pad. The wound

was sutured using wound clips (Autoclip Physicians kit, Clay Adams), which were

removed 1 week after the operation. The mice received analgesic pre‐ and postoperatively

(Temgesic 0.2 mg/kg and Rimadyl 5 mg/kg, respectively) and there covery of the animals

was followed 48 hours after surgery. After 1 week, the fat pad transplanted mice were

administered with ABT-737.

Production of retroviruses and retroviral transductions (I-III)

To produce amphotropic retrovirus particles, a retroviral construct was introduced into the

Phoenix Ampho cell line by transfecting 8 μg of DNA using JetPei transfection reagent

(Polyplus) according to manufacturers instructions. The supernatants containing

Page 69: Apoptosis by c-Myc - Helsinki

retroviruses were collected 48 hours post‐transfection and used for transduction. Prior to

retroviral infection, cells were plated at low density on dish and following day incubated

with virus supernatant. During transduction procedure the cells were centrifuged for 30

minutes 2500 rpm. Transduced cells were selected in puromycin (2.5 μg/ml), blasticidine

(10 μg/ml) or hygromycin and surviving 24 hours post-transduction.

Production of lentiviruses and lentiviral transductions (I‐IV)

The VSVg pseudotyped lentivirusparticles were produced by transfecting 293fT cells with

a three plasmid mixture consisting of the lentiviral expression or shRNA construct,

packaging construct CMVDelta8.9 and the VSVg encoding envelope construct phCMVg.

For traditional high‐titer lentivirus production, the supernatants containing lentiviruses

were collected 72 hours post‐transfection and used for transduction. Viral titer was

determined using p24 ELISA test measuring viral capsid protein p24 (Perkin Elmer,

performed in Biomedicum Genomics). Lentiviral infections of 2D cultured cells were

performed similarly as retroviral infections. Transduced MCF10A cells were selected in

blasticidine (10 μg/ml), puromycin (2.5 μg/ml) or hygromycin (250 mg/ml) 72 hours post‐transduction.

RNA isolation and quantitative real‐time PCR (qPCR) (III)

RNA was isolated by RNeasy kit (Qiagen) from cells transfected with siRNA and grown

on six-well plates for 72 h after treatment, and mRNA levels were analyzed by using the

Light Cycler 480 II instrument (Roche) using Universal Probe Library (no. 43 tggggcag)

probes (Roche Applied Science) and LightCycler 480 Probes Master mix and PCR grade

H2O (Roche Applied Science). To analyze the gene silencing activity of siRNA constructs,

the expression level of the gene of interest was compared with that of a housekeeping gene

(β-actin) and normalized to control siRNA-transduced control. Control siRNA samples

were tested in technical duplicates and the gene of interest siRNA samples in technical

triplicates. The analysis was performed by the Biomedicum Functional Genomics Unit.

SiRNA Transfections (III)

The 3′ Alexa Fluor546 siRNAs were ordered from Qiagen and transfected into the cells by

using HiPerFect transfection reagent (Qiagen). Cells were plated on 24-well plates as a

Page 70: Apoptosis by c-Myc - Helsinki

semiconfluent culture and next day transfected with siRNA. Validated siRNA sequences

are listed below.

Site-Directed Mutagenesis (III)

Expression vectors for mutagenesis were pLenti6/V5-p53_WT p53 and pLenti6/V5-p53

R175H (Bernard Futscher laboratory plasmids, Addgene: plasmids 22945 and 22936, ref.

3). Mutagenesis was performed with the QuikChange Site- Directed Mutagenesis kit

(Agilent Technologies) according to the manufacturer’s instructions. QS mutation (L22Q

and W23S) was introduced with the following primers: sense 5′-

agtcaggaaacattttcagaccaatcgaaactacttcctgaaaacaac-3′ and antisense 5′-

gttgttttcaggaagtagtttcgattggtctgaaaatgtttcctgact-3′. The presence of the mutations in the

vector was confirmed by sequencing.

Reagents (I-IV)

The following reagents were used in the experiments: 4-Hydroxytamoxifen, ABT-737 (a

kind gift from Dr Steven W Elmore, Abbott Laboratories), AP20187 (Ariad),

aminoimidazole carboxamide ribonucleotide (AICAR) (III, WB: Cell Signaling, IF:

Sigma), Etoposide, Doxorubicin, checkpoint kinase 2 (Chk2) inhibitor II (Sigma),

mitogen-activated protein kinase (MEK) inhibitor U0126 (Promega), Nutlin-3a (Alexis

Biochemicals), phosphatidylinositol- 4,5-bisphosphate 3-kinase (PI3K) inhibitor

Ly294002 (Cell Sig- naling), mitogen-activated protein kinase 1 (ERK1) inhibitor II,

ataxia telangiectasia mutated (ATM) inhibitor CGK-733 (Santa Cruz), AMPK inhibitor

Compound C, mitogen-activated protein kinase (p38) inhibitor SB203580, protein kinase,

DNA-activated (DNAk-PK) inhibitor C401, AMPK activator A769662 (Bio- Vision), and

recombinant human TNF-related apoptosis inducing ligand (TRAIL) (R&D Systems), z-

VAD-fmk (Enzyme Systems Products).

Subcellular fractionations (III)

Cells were collected into ice cold PBS and pelleted and subsequently mitochondria

enriched fraction and light membrane/cytosol fractions were collected into ice cold PBS

and isolated according to the manufactures instruction using Dounce homogenizer and

Mitochondria/Cytosol Fractionation Kit (Biovision).

Page 71: Apoptosis by c-Myc - Helsinki

Western blot analysis (I‐IV)

Cells grown on 2D cell culture were lysed either in SDS lysis buffer (II) (20% SDS in 250

mM Tris‐HCl pH 6.8 – study I) or non‐ionic Triton‐X cell lysis buffer (I) (10 mM Tris,

pH 7.5, 130 mM NaCl, 1% Triton X‐100, 10mM NaPi and 10 mM Nappi) or RIPA lysis

buffer (2 5mM Tris-HCl, pH 7.6, 150mM NaCl, 1% NP-40, 1% sodium deoxycholate,

0.1% SDS) all supplemented with EDTA‐free protease inhibitor cocktail (Roche). Mouse

mammary tumors were lysed in mammary gland (MG)‐lysis buffer supplemented with

EDTA‐free protease inhibitor cocktail (Roche). An approximately 2 mm x 2 mm piece of

snap frozen tissue was placed in tube containing MG‐lysis buffer and ceramic beads and

homogenized using Precellys homogenizer (Bertin Technologies). Subsequently, samples

were let stand on ice for 30 minutes and centrifuged at 13 000 g for 15 minutes. Protein

concentrations and western blotting was performed as described above. 10‐20 μg samples

were denatured with 5x Laemmli sample buffer and separated in 10% SDS‐PAGE gel,

and subsequently transferred to nitrocellulose filters (Bio‐Rad) followed by antigen

detection with specific antibodies and ECL™ Detection Reagents (Amersham).

Page 72: Apoptosis by c-Myc - Helsinki

RESULTS AND DISCUSSION

4. The mitochondrial apoptosis pathway is engaged by Myc in a Bak-dependent manner independently of cell context and death stimuli

Our aim was to elucidate what the molecular mechanisms are behind the Myc-induced

sensitization in epithelial cells, especially in mammary tissue. Normal or primary cells are

often resistant to drugs like TRAIL, which offers the possibility to use such molecules in

cancer therapy (LeBlanc et al., 2002). However, it is evident that not all tumors respond

well or they relapse soon after treatment. Therefore, it is important to understand which

molecular mechanism in cancer cells enables sensitivity and resistance to drugs.

Previous results show that Myc activation for instance in growth factor deprived cells or

drug/DNA damage stressed cells results in apoptosis in a mitochondrial-dependent manner

(Klefstrom et al., 2002). Likewise, Myc seems to sensitize cells to variety of different

stresses, however the exact mechanism behind mitochondrial pathway engagement

remains diverse. In our study, we analyzed how mammary epithelial cells respond to

ectopic expression of Myc by using MYC-ERtm cells (Eilers et al., 1989; Littlewood et al.,

1995). We show that activation of Myc does not induce apoptosis in normal, two-

dimensional cell culture as shown previously in other non-transformed cell types however,

Myc can prime mitochondria for various drugs (TRAIL and Etoposide) by engaging the

mitochondrial apoptosis pathway (I-III). Furthermore, Juin et al, showed that Myc

activation results in cytochrome c release in Rat-1 cells (Juin et al., 1999). In accordance,

we showed that combined Myc and TRAIL signaling results in cytochrome c release into

the cytosol, however, Myc alone is unable to induce MOMP (I-III). Previous results

suggest that Bax is the major mediator of Myc-mediated apoptosis in primary mouse cells,

pancreatic β-cells, and in Εμ−Myc lymphoma (Annis et al., 2005; Dansen et al., 2006;

Eischen et al., 2001a). Concordantly, our results show that Bax undergoes a

conformational change and its N-terminus is exposed in cells undergoing apoptosis by

Myc and TRAIL. In addition, we show also that Myc and TRAIL activation was able to

induce also Bak activation and N-terminus exposure in apoptotic cells (I and III).

Importantly, this was the first time when Bak was suggested to mediate Myc functions.

Furthermore, we were able to show using RNAi-techniques that Bak is required for Myc-

mediated engagement of mitochondrial pathway in mammary epithelial cells (I and III). In

Page 73: Apoptosis by c-Myc - Helsinki

contrast, similar analysis of Bax-deficient cells showed that Bax is not necessarily

required for Myc apoptosis in these cells. Surprisingly, an evident Bak up-regulation was

noticed in Bax-deficient cells indicating compensational mechanisms could be involved

(our unpublished results) (Mondal et al., 2012). Also, in variety of models tested

abrogation of apoptosis often requires Bax, Bak double deficiency (Wei et al., 2001).

Furthermore, we showed that Myc was not able to induce transcriptional changes in

protein expression levels on Bad, Bax, Bcl-xL, Bid, Bim or Puma as suggested earlier (I

and our unpublished results) (Hemann et al., 2005; Maclean et al., 2003; Mitchell et al.,

2000). In contrast, a transient Bak upregulation was seen at 10-hour time point in Myc-

activated cells (I). However, whether this is a direct transcriptional target by Myc remains

to be elucidated. These results suggest that Myc-induced mitochondrial priming occurs via

the Bcl-2 family members and in a Bak-dependent manner in various cell types: fibroblast,

mammary epithelial cells and variety of other epithelial cells, cancer cell lines, and

primary mouse mammary cells (I-IV).

In addition, to explore how different physiological environments and epithelial integrity

can affect the capability of Myc to induce proliferation and apoptosis we used a three-

dimensional organotypic cell culture model of mammary epithelial cells (Debnath et al.,

2002). In this system, epithelial cells start to proliferate, undergo apoptosis to form a

hollow lumen and possibly migrate to form a rounded, organized, polarized and quiescent

structure called acini. Our results show that Myc activation can induce apoptosis if

activated during early acinar morphogenesis when the cells are unorganized and

proliferating (II). In contrast, fully-formed acini were resistant to Myc activation and

apoptosis and also to death induction by TRAIL alone suggesting that either polarity or

quiescence could prevent Myc functions (II). In accordance, Zhan et al. showed that Myc

functions are suppressed by cell polarity and deregulation of Scribble can promote Myc-

driven mammary tumorigenesis (Zhan et al., 2008). Importantly, activation of Myc in

fully formed acini was able to sensitize quiescent cells to TRAIL indicating that Myc is

still active and can augment apoptosis. More importantly, the proliferative and apoptotic

capacity of Myc seemed to be different in acini at different morphogenic stages. For

instance, if the cells were non-polarized and unorganized at the time of Myc activation,

like in early acinar morphogenesis, activation of Myc resulted in bigger acini. Whereas, in

fully-formed acini Myc was unable to re-initiate the cell cycle (II). These results indicate

Page 74: Apoptosis by c-Myc - Helsinki

that organized epithelial architecture or proliferation driven by acinar morphogenesis is

critical for tumorigenic functions of Myc, however apoptotic sensitivity remains

independently of proliferative status. In this study we also showed that loss of LKB1

inhibited polarization and organization of acini, as measured by GM130, β-catenin and

α6-integrin (II). However, these LKB1-lacking acini did finally undergo quiescence, and

interestingly, Myc was then able to induce cell cycle re-entry. Interestingly, our results

indicate that Lkb1 inhibits tumorigenic functions of Myc by maintaining the polarity of

epithelial cells in acini. Furthermore, in addition to its polarity function, the Lkb1 protein

has been shown to regulate proliferation by inducing growth arrest, and also activation of

metabolic pathways via AMPK (Tiainen et al., 1999). Thus the results indicate that

organized epithelial architecture is one of the critical suppressors of proliferative Myc

function. Furthermore, acinar morphogenic program and differences in the proliferation

status of acini may impact on the cell, for instance, in metabolism. Hence, in this acini

model, the exact role of active cell cycle machinery and differential activation of

metabolic pathways on Myc function remains to be elucidated. Moreover, our result

indicated that Myc also engaged mitochondrial apoptosis pathway in the three-

dimensional acini model since Bim- and Bid-deficient cells were resistant to Myc and

TRAIL synergistic death. Interestingly, Bak seem to be required for Myc-apoptosis in

early acinar structures (II and our personal data).

Finally, the role of mitochondrial apoptosis pathway in Myc-apoptosis was emphasized

since it was inhibited by anti-apoptotic Bcl-xL. Likewise, previous results suggest that Bcl-

xL and Bcl-2 are key regulators of Myc tumorigenesis and apoptosis (Eischen et al.,

2001b). Furthermore, the BH3 mimetic ABT-737, which inhibits anti-apoptotic Bcl-2

family members, Bcl-2, Bcl-xL and Bcl-w, phenocopied the effect of Myc activation and

sensitized cells to apoptosis (I and II). Intriguingly, since the effect of ABT-737

significantly resembled apoptotic activation by Myc, it could be hypothesized that Myc

influences the balance of anti-apoptotic Bcl-2 family members. Altogether, these results

indicate that Myc-mediated priming of apoptosis is mitochondria-dependent.

Page 75: Apoptosis by c-Myc - Helsinki

5. The mitochondrial amplification loop as a mechanism for lethal

caspase activation in death receptor signaling

Previous studies have implied that during TRAIL death receptor signaling, procaspase-8 is

processed and activated at the DISC after ligand binding and in type II cells this apoptotic

signal is directed to the mitochondrial apoptosis pathway by Bid protein (Broaddus et al.,

2005; Kischkel et al., 2000). To understand the apoptotic co-operation of TRAIL and Myc

signaling, initiator caspase-8 and Bid were silenced using RNAi. Indeed, our results

indicate that both caspase-8 and Bid are required for TRAIL signaling (I-II). Furthermore,

our results demonstrate that TRAIL induces caspase-8 activation in these TRAIL-resistant

mammary epithelial cells and fibroblasts, however evident Bid cleavage was seen only in

cells where Myc was also active (I).

In apoptotic cells by Myc and TRAIL signaling, an interesting augmentation of caspase-8

cleavage was observed when the mitochondrial apoptosis pathway was activated. This

enhanced caspase-8 processing and activity was inhibited in cells if mitochondrial

pathway was blocked (I). Previously such post-mitochondrial Bid and caspase-8

processing has been suggested to occur downstream of the mitochondria, possibly by

caspase-3 (Slee et al., 2000).

FLIP is a caspase-8 inhibitor recruited to the death-inducing signaling complex (DISC)

instead of caspase-8. It has been suggested that Myc would directly increase FLIP

expression levels and thereby increase sensitivity of cells to TRAIL (Ricci et al., 2004b).

However, since Myc sensitizes cells to a variety of death stimuli UV, genotoxic drugs and,

for example, to growth factor deprivation, it is highly possibly that FLIP alone does not

explain apoptotic sensitivity (Askew et al., 1991; Juin et al., 1999; Prendergast, 1999). In

addition, we were not able to see any significant mRNA changes of FLIP in gene

expression profiling of Myc-activated cells (our personal results). Therefore, to answer

this question we analyzed Myc-induced caspase processing by using dimerizable caspase-

8 constructs that are independent of DISC formation. We showed that combined

expression of Myc and dimerized caspase-8 was able to induce cell death (I). These results

indicate that mechanisms other than enhanced DISC-recruitment explain synergistic cell

death between Myc and TRAIL, and results in a mitochondrial amplification of caspase

Page 76: Apoptosis by c-Myc - Helsinki

signaling. Furthermore, mitochondrial pathway-specific ABT-737 was also able to

augment caspase-8 processing and activation when combined with TRAIL (I), also

suggesting a mechanism downstream of DISC for Myc.

6. Myc-induced Bak N-terminal exposure is a conformational change

preceding Bax activation and MOMP

Pro-apoptotic Bak normally exists embedded in the mitochondrial outer membrane in a

conformation where its N-terminus and BH domain are buried (Dewson et al., 2008).

Instead, Bax is more loosely attached to mitochondrial membranes and can be translocated

to the mitochondria upon death stimuli in some cells (Schellenberg et al., 2013). When the

mitochondrial apoptosis pathway is activated, both of these proteins are thought to

undergo a conformational change resulting in exposure of the N-terminus and transient

exposure of the BH3 domain, finally enabling a homo-oligomerization of these proteins

(Annis et al., 2005; Dewson et al., 2009). How oligomerization really occurs, and whether

Bak and Bax molecules are interacting with each other and how the exact pore is formed

is still under depate. In our experiments, when Myc was activated in mammary epithelial

cells, N-terminal exposure of Bak was detected in immunofluorescence staining using

non-detergent conditions (I and III). This was clearly different from Bax activation that

occurred when cells were in addition treated with TRAIL and underwent apoptosis.

Importantly, this “preactivation” of Bak did not coincide with Bak homo-oligomerization

and proper oligomers were visible only during apoptosis. Likewise, Moldoveanu et al.

suggest that in Bak activation process and conformational changes, it is Bid that launches

the dimerization and MOMP-inducing oligomerization of Bak (Moldoveanu et al., 2013)

Furthermore, in our analysis Bak and Bax were localized to mitochondria both in control

and Myc-expressing mammary epithelial cells, however Bax was also found from cytosol

(III and our personal results). This indicated that Myc does not enhance Bax translocation,

on the other hand it still might affect on Bax activity as suggested earlier (Soucie et al.,

2001)

According to the neutralization model, sequestration of Bak binding inhibitors would be

enough to enable Bak conformational change (Tait and Green, 2010). Hence, to further

Page 77: Apoptosis by c-Myc - Helsinki

understand how Bak activation is regulated, we analyzed the role of Bak interacting anti-

apoptotic proteins Bcl-xL, Mcl-1, and A1 using RNAi (Willis et al., 2005). Surprisingly,

only Bcl-xL deficiency resulted in Bak conformational change and activation. In addition,

sequestration of Bcl-xL, Bcl-2, and Bcl-w with ABT-737 resulted in Bak activation (III).

This indicated that Bak acitivity preceeding MOMP is regulated by BclxL-interactions.

However, our interaction analysis showed no differences between Bak-Bcl-xL and Bak-

Mcl-1L in Myc-expressing cells, whereas inhibition of Bcl-xL by ABT-737 resulted in loss

of interaction and both are able to activate Bak (III). Importantly, Bcl-xL, overexpression

in mammary cells was still able to abrogate Myc-induced Bak activation and apoptosis,

suggesting that Bcl-xL might have still have inhibiting role in Myc engagement of

mitochondria (I and III). Indeed, in Myc and TRAIL-treated apoptotic cells this Bcl-xL

association to Bak was diminished almost completely, indicating that Bcl-xL regulates

conformational changes and oligomerization of Bak (our personal results). According to

our personal results, the role of Mcl-1 seems to be evident in TRAIL-signaling pathway

but not important regulator of Myc functions, since we could detect Mcl-1 cleavage and

loss of Bak-Mcl-1 interaction in apoptotic cells by Myc and TRAIL as suggested

previously (Weng et al., 2005). In accordance, our results suggest that Mcl-1

overexpression is able to slightly inhibit cell death by Myc and TRAIL, but not Bak

activation induced by Myc (our personal results). These results support the “neutralization

model” for Bak activity regulation, where Bcl-xL-free Bak undergoes conformational

change and exposes its N-terminus. Conversely, Myc did not induce such Bak liberation

from Bcl-xL suggesting a different mechanism. Hence, according to another ”direct

activation” model Bak activation and oligomerization always requires a another direct

activator BH3-only protein like TRAIL-tBid (Tait and Green, 2010). Likewise, our results

indicate that Bak oligomerization and MOMP require additional activator. Furthermore,

another transient conformational change, BH3 domain exposure, is suggested to follow N-

terminal exposure and occurs before final oligomerization of Bak and can also be involved

in Myc priming (Dewson et al., 2009). Moreover, mitochondrial fission during apoptosis

has major role in regulating MOMP (Suen et al., 2008). Thus, it is possible that Myc-

induced changes at mitochondrial structure and fission can affect interactions between the

Bcl-2 family members and, more importantly, Bak activation. Altogether, these results

suggest a hypothetical model where these two models are combined, and highlights that

Page 78: Apoptosis by c-Myc - Helsinki

N-terminal exposure of Bak is a separate step from final activation, which also includes

BH3 domain exposure and homo-oligomerization.

7. BH3 mimetic reactivates mitochondrial apoptosis in Myc mammary

tumors Myc is highly over-expressed in breast cancer, for example, via amplification (Naidu et al.,

2002). Furthermore, Myc-induced tumorigenesis often requires secondary hits such as

Bcl-xL overexpression, or mutations in Ras or p53, which enable Myc to fully induce

proliferation since its apoptotic phenotype is inhibited (Dang, 2012; Kelly et al., 2011).

However, it is possible that there are apoptotic pathways in Myc overexpressing tumor

cells that could be easily re-activated by specific inhibitors or activators, making these

secondary mutations defective. ABT-737 is a BH3 mimetic small molecule inhibitor that

was originally designed to act like a de-repressor of the anti-apoptotic Bcl-2 family

proteins, Bcl-xL, Bcl-2 and Bcl-w (Oltersdorf et al., 2005). Our previous results from Bak

activation studies suggested that key regulators for Myc-induced mitochondrial priming

were Bak and proteins regulating Bak (I, III). Hence, it was not surprising that WAP-Myc

mouse mammary tumors were sensitive to ABT-737. Likewise, human mammary

epithelial cells expressing Myc synergistically induced apoptosis with ABT-737 (our

unpublished results). Furthermore, our analysis showed that Wap-Myc tumors expressed

Bcl-xL (IV). Therefore, we wanted to explore whether BH3 mimetic ABT-737 would be

such re-activator of Myc apoptosis in WAP-Myc mammary tumors. In our experiments

Wap-Myc tumors arose spontaneously with a 3.5-month latency and subsequent

administration of ABT-737 was able to inhibit tumor growth (IV). In accordance, Kelly et

al. also showed that ABT-737 augments apoptosis of preneoplastic B-lymphoid cells of

Eμ-myc transgenic mice and prolonged lymphoma-free survival (Kelly et al., 2011). The

amount of individuals with presence of pulmonary metastases was diminished to half in

ABT-737-treated animals. Importantly, the endogenous Wap-Myc tumors were

heterogeneous in growth rate and probably also with genetic backgrounds, all acquiring

different secondary mutations. Previous results show evidence that resistance to ABT-737

is associated with overexpression of Mcl-1 (Konopleva et al., 2006; van Delft et al., 2006).

However, the survival and growth rates of ABT-73-treated Myc-tumors did not show any

correlation with Bcl-2 family protein expression levels when compared to control group.

Page 79: Apoptosis by c-Myc - Helsinki

Further analysis could reveal, whether ABT-737 induces evolution of tumors involving

changes in Bcl-2 family proteins expression levels. Thereby, tumor might acquire

resistance. Subsequently, another approach was taken and Wap-Myc tumors were

syngrafted. Importantly, the syngrafted tumor resembled endogenous Myc tumor and were

positive for basal marker (CK14) and epithelial cell markers and had basement membrane

disorganized. Similarly, ABT-737 administered syngrafted Myc tumors grew slower and

tumors showed markers for increased apoptosis, however no differences were seen in

survival. ABT-737 was apparently toxic for mice that had undergone a surgery and had

wound still healing at the place of intraperitoneal injections given daily. This resulted in

an unexpected death of ABT-737 treated mice. Of note, one of the side-effects of

treatment with ABT-737 is acute thrombocytopenia (low platelet count) (Wilson et al.,

2010). Thus, the results suggest, that ABT-737 might re-activate apoptosis in Myc tumors,

however, to be able to see an effect on survival, a less toxic orally bioavailable ABT-263

should be used.

8. Cytoplasmic tumor suppressor p53 is activated by Myc to prime

mitochondrial apoptosis Emerging evidence suggests that p53 is involved in Myc-induced tumorigenesis and

apoptosis (Hemann et al., 2005). In addition to growth arrest, p53 has major role in

mediating the mitochondrial apoptosis pathway after DNA damage, for example, by

transcribing Bcl-2 family members Puma and Bax (Chipuk et al., 2005). Furthermore

when inducing apoptosis, p53 can act independently of its nuclear function and directly

interact and activate Bak and/or Bax at mitochondria and, can be inhibited by Bcl-xL

(Eischen et al., 2001a). Surprisingly, in our analysis p53 was found from mitochondrial

and cytosolic compartments in mammary epithelial cells if Myc was activated (III). In

similar experiments, drug-induced DNA damage was able to stabilize p53 at the nucleus

and induce upregulation of growth inhibitory genes like p21 indicating a different

mechanism for p53 action. Concordantly, neither could the DNA damage induce Bak

activation what Myc did. Previously, activation of oncogene Myc and Myc-induced DNA

damage have been shown to result in phosphorylation of Serines 15 and 37 of p53.

Importantly, phosphorylation seems to be required for inhibition of Myc-induced

Page 80: Apoptosis by c-Myc - Helsinki

tumorigenesis (Pusapati et al., 2006; Sluss et al., 2010). Also our results show that Myc

stabilizes p53 and phosphorylates p53 at Serine 15 in various cell types and also Wap-

Myc mouse mammary hyperplastic tissue expressed Ser15 phoshorylated p53 (III).

Interestingly, a small molecule Mdm2 inhibitor, Nutlin-3a, stabilized and accumulated p53

both at nucleus and mitochondria, and also activated Bak. In contrast, Nutlin-3a was able

to able induce p53-mediated transcription of p53 target genes, like TIGAR, whereas Myc

was not. Likewise, Mdm2 inhibitor ARF has been shown to be involved in Myc-induced

apoptosis, however whether ARF in involved in mitochondrial function of p53 remains to

be elucidated (Zindy et al., 1998). All together, this data does not exclude that

phosphorylation of p53 in Myc-activated cells would not involve DNA damage, but is

suggests that drug-induced DNA damage and Myc activation direct p53 at different

compartments at cell and possibly leading to different phenotypes. In addition, ectopic

transcription-deficient p53QS-mutant experiments in p53-null MCF10A cells indicated that

p53-mediated Bak activation was a transcription-independent phenomenon (III). Thus it

was evident that p53 seemed to have a cytoplasmic role in Myc-induced mitochondrial

priming and Bak activation.

The Bak and Bax are regulated either by inhibiting anti-apoptotic interaction, or directly

activating BH3-only protein (Tait and Green, 2010). Importantly, conformational

activation of Bak is shown to be regulated by Bcl-xL and Mcl-1-mediated BH3:groove

interactions (Willis et al., 2005). However in our analysis, Myc activation did not change

Bcl-xL or Mcl-1 association to Bak as measured by co-immunoprecipitation (III).

Interestingly, ABT-737 has been shown to also release Bcl-xL from p53 and indeed, we

saw loss of p53 and Bcl-xL interaction in ABT-737-treated epithelial cells. Previously, p53

has been suggested to bind the “underside” of the Bcl-xL and therefore p53 and Bak

should not compete for the same interaction site, the hydrophobic binding groove (Hagn et

al., 2010). Surprisingly, Myc activation also decreased Bcl-xL association with p53,

possibly indicating free apoptotic p53 that could thereafter be able to directly activate Bak

as suggested earlier (Chipuk et al., 2004). Additionally, interaction analysis using a

proximity-ligation assay of p53 showed that Myc activation results in co-localization of

p53 and Bak, and, a simultaneous decrease in interaction of p53 and Bcl-xL (III). However,

to completely understand the role of Bcl-xL-Bak interaction other approaches should be

considered. In addition, post-translational modification of Bcl-xL has been indicated to

Page 81: Apoptosis by c-Myc - Helsinki

affect the function and interactions of Bcl-xL and could be involved in Bak activation by

Myc (Basu and Haldar, 2003). However, we did not see, for example, phosphorylation of

Serine 62 of Bcl-xL in Myc-activated cells (our personal results). Finally, results from

p53-deficient cells demonstrate that Myc-induced Bak activation is p53-dependent (III).

Furthermore, our RNAi screen to decipher Bcl-2 family members involved in Myc and

TRAIL apoptosis suggested that p53 interaction partner and derepressor BH3-only protein

Puma could be involved in Myc and TRAIL synergistic death (our personal results).

However, the exact role of Puma in regulating Bak activation or p53 functions by Myc

need to be further elucidated. Thus these data indicate that phosphorylated p53 in Myc

expressing cells is Bcl-xL free and accumulates at the mitochondria to possibly directly

mediate activation of Bak.

9. The energy sensor AMPK is behind the Myc-mediated apoptotic

sensitivity (III)

P53 is phosphorylated and stabilized by a variety of kinases involved in the DNA damage

response. Altogether p53 has 15 sites for phosphorylation some increasing tetramerization,

some decreasing Mdm2 association (Ser15) and a few improving acetylation (MacLaine

and Hupp, 2011). The p53 Serine 15 has been suggested to be crucial for the apoptotic

function of Myc in mouse models (Jones et al., 2005). In our analysis, pharmacological

screening of kinases mediating Myc-induced phosphorylation of p53 revealed a tumor

suppressor AMPK as one of the possible kinases involved (III). AMPK is known to be

involved in energy stress-mediated growth inhibition and phosphorylates p53 during stress

(Jones et al., 2005). In addition, AMPK is known to inhibit catabolic pathways (glycolysis,

fatty acid oxidation and mitochondrial biogenesis) and inhibits anabolic pathways

(gluconeogenesis, glycogen and fatty acid and protein synthesis) (Luo et al., 2010). In

contrast, less data exists about the role of AMPK in mediating apoptosis. Furthermore, our

results demonstrate that pharmacological activation of AMPK was able to induce

phosphorylation of p53, and subsequent Bak activation and cell death sensitivity (III).

Interestingly, AMPKT172 and its phosphorylation target ACCSer79 were increasingly

phosphorylated in Myc-expressing cells and Wap-Myc mouse mammary hyperplasia,

indicating AMPK activation as well. Previously, Lkb1 kinase, CAMKK, and TAK have

Page 82: Apoptosis by c-Myc - Helsinki

been suggested to phosphorylate and activate AMPK, however it remains to be elucidated

in the future through which kinase AMPK is phosphorylated in Myc-expressing. However,

based on our results Lkb1 has tumor suppressive function against tumorigenic Myc. Taken

together, the role of AMPK activation in mediating apoptosis seems to be evident. These

results suggest that AMPK is a key protein in mediating activation of p53 in Myc

expressing cells.

10. Bak activation coupling oncogene-mediated metabolic stress and

apoptosis

AMPK activation is an energy sensor mechanism that responds to the increase of AMP or

decrease in ADP levels in the cells (Hardie, 2011). Interestingly, our data suggests that

ATP levels are dropping and ADP/ATP ratio increases (III).

Importantly, AMPK has an emerging role as a nutrient and energy sensor that maintains

energy homeostasis by increasing catabolic pathways and simultaneously acting as tumor

suppressor (Hardie et al., 2012; Luo et al., 2010). Therefore, it was speculated that

AMPK was activated as a response to such metabolic stress induced by Myc. Previous

data show that Myc stimulates glutamine uptake and metabolism through up-regulation of

glutamine transporters and glutaminase (GLS) (Wise et al., 2008). Likewise, in our

analysis Myc was able to up-regulate GLS expression and rendered cell addicted to

glutamine metabolism (III). Furthermore, these data raises a possibility that AMPK is

involved in Myc-induced metabolic reprogramming, which can benefit tumorigenesis

(Liang and Mills, 2013). In accordance with previous publications, Myc was also able to

induce glycolysis in mammary epithelial cells (our personal results). Additionally, nutrient

availability seems to be a determining factor for Myc-induced metabolic pathway

activation since, for instance, under glutamine-deprivation Myc was shown to direct

reprogramming towards glycolysis (our personal results). Taken together, all these

changes are indicative for activation of catabolic pathways and oncogene-induced

metabolic transformation suggesting that Myc-expressing cells are simultaneously

undergoing energy stress and metabolic reprogramming.

Furthermore, Myc, AMPK, or p53 activation, were able to induce Bak conformational

change and sensitize mitochondria to apoptosis (I-IV). Since Myc-induced metabolic

Page 83: Apoptosis by c-Myc - Helsinki

reprogramming was co-incidental with AMPK and Bak activation, it could be

hypothesized that Bak activation is a result of a metabolic stress. Interestingly, breast

cancer cells expressing Myc had an increased level of Bak activation in

immunofluorescence analysis (III), indicating a possible oncogene stress indicated by Bak

N-terminal exposure. Previous results suggest that “oncogene stress” is partly results from

the enforced induction of cell cycle entry (mitotic stress or replicative stress) by oncogene

that makes cells more prone to apoptosis (ref). However, data presented here indicate that

oncogene-induced metabolic stress is not necessarily the result of excessive proliferation,

since Myc activation in fully-formed mammary acini was able to induce apoptotic

sensitivity without increase in proliferation status. Furthermore, neither did Myc activation

in two-dimensional cell culture induce any evident increase in cell population, even it did

induce cell cycle re-entry in growth deprived conditions (II and our personal results).

Importantly, Myc expressing cells had increased catabolic pathway activity, such as

glutaminolysis, fatty acid oxidation and glycolysis, indicating metabolic reprogramming

(Ward and Thompson, 2012). Hence, a more plausible explanation for stress in Myc

overexpressing cells could be the activation of multiple signaling pathways at the same

time and energy-stress deriving from the usage of catabolic pathways and overall increase

in biosynthesis (Dang, 2013). Once these energy-stressed cells are declined with nutrients,

they may not be able to cope with crucial tasks and become more vulnerable to apoptosis.

Already, some discussion has emerged about ideas how metabolism could impinge on

apoptotic signaling pathway and those involve p53, caspase-2 or Bcl-2 family members

(Buchakjian and Kornbluth, 2010). For instance, growth arrest and metabolism have

already been previously linked via AMPK and p53 (Buchakjian and Kornbluth, 2010).

Furthermore, Bcl-2 family members have shown to directly regulate metabolic pathways

via Bad acting as glucokinase in glycolysis and via Noxa-Mcl-1 interaction balancing

apoptotic sensitivity during glucose insufficiency (Alves et al., 2006; Danial et al., 2003).

In addition, our results indicate that nutrient availability can affect on metabolic

reprogramming and apoptotic sensitivity (III, unpublished results). Importantly, as

mediators of apoptotic sensitivity AMPK and p53 are acting as tumor suppressors in Myc-

expressing cells, however their function as regulators metabolic reprogramming is

possible. All together, these data suggest that the metabolic and mitochondrial apoptotic

pathways can be coupled in a novel AMPK-p53-Bak-dependent way (Figure 13).

Page 84: Apoptosis by c-Myc - Helsinki

Figure 13. Model for Myc-effects in mammary cells. Myc-induced activation of signaling

pathways involved in cell growth and proliferation requires metabolic adaptation. This metabolic

reprogramming enables growth, however, it also lowers the cellular energy as measured by ATP

levels, which is recognized by the energy sensor AMPK. Consequently, AMPK phosphorylates the

tumor suppressor p53 and, thereby p53 is stabilized and accumulating at the mitochondrial

compartment. Finally, mitochondrial p53 engages and activates Bak that makes mitochondrial

apoptosis pathway primed for apoptotic stimuli. Furthermore, this model indicates that by targeting

metabolic pathways or Bcl-2 family members in Myc-expressing tumor cells, cell can become

vulnerable to death. All together, this model explains how Myc-induced metabolic reprogramming

and cell death sensitivity can be coupled.

Page 85: Apoptosis by c-Myc - Helsinki

CONCLUSIONS The studies presented here provide evidence that oncogene Myc-induced apoptosis is

mitochondrial dependent in mammary epithelial cells and in other cell types. Our findings

expose new evidence for Bcl-2 family member mediated control of Myc-induced

apoptosis and Bak-dependent priming of the mitochondrial pathway by a mechanism

involving mitochondrial p53 (I-III). Furthermore, data here suggest a novel definition for

Bak activation and N-terminal exposure as a conformational change that is separate from

BH3 exposure and oligomerization of Bak. We suggest that the conformational change of

Bak is an apoptotic priming mechanism whereas the oligomerization of Bak represents the

pore forming activity, considered as a point-of-no-return in the apoptotic process. Hence

we open new ideas for the current “direct activator/derepression” model suggested earlier

(Tait and Green, 2010). The data presented herein also suggest that BH3 mimetic drugs

can sensitize tumor cells to Myc-apoptosis in vivo (IV), which also highlights the role of

Bcl-xL as a key controller of Myc-dependent apoptosis in epithelial tumors and points new

therapeutic possibilities to treat breast cancer.

Emerging data indicate that p53 has transcriptionally independent functions and that

cytoplasmic p53 has apoptotic function involving interactions with Bcl-xL and Bak/Bax

(Chipuk et al., 2004; Sot et al., 2007). Correspondingly, our results indicate that p53

cytoplasmic p53 has a crucial role in Myc-induced priming of the mitochondrial apoptosis

pathway (III). According to our results, p53 is required for Myc-induced engagement of

the mitochondrial apoptosis pathway via transcription-independent mechanisms, by

functioning as a direct activator for Bak (III). Therefore we suggest a Bak activation

model, where Bak is kept in check and the N-terminus hidden by Bcl-xL. Myc or the

pharmacological inhibition of Bcl-xL is enough for the Bak to be able to undergo the first

conformational change, exposure of the N-terminus. In addition, our results indicate that

Myc induced activation of the AMPK and phosphorylation of its downstream target p53.

Furthermore, we hypothesized that the phosphorylated p53 accumulates in the

mitochondria and by altering the interaction between Bak and Bcl-xL, the Bak finally

becomes activated. Moreover, it could be speculated that Myc activation could induce a

partial release of Bak from tight inhibition by Bcl-xL where Bcl-xL is adopting a different,

yet unknown, conformation allowing the first Bak conformational change. However, more

Page 86: Apoptosis by c-Myc - Helsinki

studies are needed to clarify how these molecular level alterations lead to conformational

activation of Bak.

Finally, the results reveal that Myc-induced metabolic reprogramming and apoptotic

sensitivity are tightly coupled. Here we suggest that activation of Myc promotes metabolic

transformation by inducing glycolysis, activating the glutaminolytic pathway (GLS), and

by promoting a catabolic mode of metabolism through inhibiting fatty acid synthesis (III).

Furthermore, Myc-induced metabolic transformation decreased cellular ATP levels,

activating the energy sensor AMPK that releases carbon for bioenergetic reactions and

promotes biosynthesis of nucleic acids. Thus, we show that Myc-induced metabolic stress

and direct activation of AMPK both induce apoptotic sensitivity via p53 and Bak-

mediated pathway. However, future studies are still needed to understand how, for

example, Myc-induced glutamine-dependence (addiction) is linked to the

apoptotic mechanisms described here or how p53 mechanistically activates Bak at the

surface of the mitochondria. Interestingly, these results also indicate that nutrients and

their availability and sensing affect on transformation capacities and apoptotic sensitivity

of cancer cells possibly in a very critical manner and should be considered carefully in

cancer therapies. The contextual role of AMPK as part of the cancel cells metabolic

reprogramming and as a tumor suppressor was emphasized by our results and will be

important question in future. Furthermore, it will be interesting to explore possibilities to

use drugs that modulate AMPK activity as therapeutic sensitizers of tumor cells to Myc-

dependent apoptosis. Finally, our results bring new insight into oncogenic stress and

apoptotic sensitivity by Myc highlighting the role of altered cell metabolism (1-IV). All

together, it will be future challenge to understand how that knowledge of crosstalk

between metabolic stress and apoptotic sensitivity can be used for therapeutic

interventions in cancer.

Page 87: Apoptosis by c-Myc - Helsinki

ACKNOWLEDGEMENTS

The work presented in this dissertation has been carried out at the University of Helsinki in the department of Biochemistry and Developmental biology at the Institute of Biomedicine in Biomedicum Helsinki and under the Translational Cancer Biology Research Program. I would like to express my gratitude to professors Esa Korpi, Olli Jänne, Sampsa Hautaniemi, Lauri Aaltonen, Hannu Sariola and Tomi Mäkelä for providing the excellent working facilities. I am most grateful to my supervisor, Juha Klefström, for all his support and giving me the possibility to start my science career and to work in his group. More importantly, Juha is acknowledged for giving me the opportunity to take responsibility of my own research goals and independently design experiments to test hypothesis. Juha is also thanked for being an example of enthusiastic scientist, always creating an interesting story. I sincerely acknowledge Professors John Eriksson and Ville Hietakangas for their grateful review of the thesis on a tight schedule. Minna Ahvenainen is warmly acknowledged for her revision of the English language. John Eriksson and Philippe Juin are also acknowledged for their guidance and valuable comments as members of the thesis follow-up committee. Also Professor John Eriksson and Tapio Heino are thanked for reviewing my master thesis. I wish to thank also Helsinki Biomedical Graduate School for all financial support. I would like to warmly thank all my collaborators especially Tomi Mäkelä and Outi Monni for their valuable comment and support. Also Panu Kovanen is acknowledged for expertise in histopathology and Daniel Laajala and Tero Aittokallio for their knowledge and input. This study would not have been possible without all the wonderful people, friends and colleagues, around me. I want to thank all the present and former Klefalab members that participated in the lab meetings with intelligent comments and made our working days enjoyable. Especially I like to acknowledge all co-authors: Vilja Eskelinen, Johanna Partanen, Annika Hau, Topi Tervonen, Sirkku Saarikoski, Heidi Haikala, Elsa Marques, for excellent input. Special thanks go to Johanna Englund and Annika Hau who shared with me the early days in the lab and Vilja Eskelinen for being nice and reliable co-worker, you were always there for me if needed. Also borrowed postdoc Niklas Ekman is thanked for sharing knowledge and opinions about everything with us. My deepest gratitude goes to very helpful technical stuff from Klefalab and Biomedicum Genomics: Tiina Neejärvi, Tarja Välimäki, Teija Inkinen and also secretary Anu Taulio is thanked for everything. Many things would have gone very difficult without you guys. Tiina, Vilja and Heidi are especially thanked for enormous helping during my leaves and enabling working from home.

Page 88: Apoptosis by c-Myc - Helsinki

Especially, Professores Tomi Mäkelä and Päivi Ojala are acknowledged for constructive comments during lab meetings and retreats. Also their lab members, specifically Katja, Emma, Kari, Suski, Annika, Pekka, Anou, Liina, Eeva, Tea, Thomas and their helpful technicians Kirsi and Birgitta are thanked for crateful discussion and most importantly, for all the fun we had. Special thanks go to Kari for important comments concerning AMPK. I also want to thanks my colleagues Essi Havula, Emma Kuuluvainen, Anne Mäkelä, Katja Helenius and Vilja Eskelinen for sharing science and life with me. Also all people attending Kalakisat and Ahvis are thanked for the fun and unforgettable memories. My colleagues Maria Rajecki, Henna Heinonen, Minna Jääskeläinen, Mikko Turunen and Tatyana Lepikhova are thanked for giving me the opportunity to be involved in their interesting projects, it has been pleasure. I also want to thanks my colleagues and study friends from Viikki times for fruitful discussions about working in the science world and being always so positive. Special thanks goes to my dearest friends Salla, Janica, Anski and Sointu. Your support and company has been invaluable. It is always relaxing to discuss with you about something else than science, for example, teaching children… Most importantly, you have walked with me during my whole life, listening, teaching, traveling and having fun. Finally, I wish to thank my family for love and support, especially my husband Tero for his patience and for keeping our life busy with football. Most of all, my son Onni and daughter Emmi, are sincerely thanked for teaching me a lot in life – you are my sunshines. I also want to express my gratitude to my parents, Jouko and Marjut, for teaching me to question things in life and not to depend on other people`s opinions and for always encouraging me for challenges. I also would not be here without my wonderful sisters, Elina and Minna, and brothers, Köpi and Aaro, who always are there for me and were in huge help with combining my thesis and children. Also Tero`s parents and my grandparents, Aili and Eino, are thanked for endless care and support for thesis. This work was supported by personal grants rom the Academy of Finland, the Sigrid Juselius Foundation, the Finnish Cancer Organizations, the Biomedicum Foundation, the Emil Aaltonen Foundation, the Ida Montini Foundation, the Orion-Farmos Research Foundation, the Paulo Foundation, University of Helsinki and Helsinki Biomedicum Graduate Program. Helsinki, March 2014

Page 89: Apoptosis by c-Myc - Helsinki

REFERENCES Adams, J. M., Harris, A. W., Pinkert, C. A., Corcoran, L. M., Alexander, W. S., Cory, S., Palmiter, R. D., and Brinster, R. L. (1985). The c-myc oncogene driven by immunoglobulin enhancers induces lymphoid malignancy in transgenic mice. Nature 318, 533-538. Adhikary, S., and Eilers, M. (2005). Transcriptional regulation and transformation by Myc proteins. Nature reviews Molecular cell biology 6, 635-645. Ahuja, P., Zhao, P., Angelis, E., Ruan, H., Korge, P., Olson, A., Wang, Y., Jin, E. S., Jeffrey, F. M., Portman, M., and Maclellan, W. R. (2010). Myc controls transcriptional regulation of cardiac metabolism and mitochondrial biogenesis in response to pathological stress in mice. J Clin Invest 120, 1494-1505. Alitalo, K., Bishop, J. M., Smith, D. H., Chen, E. Y., Colby, W. W., and Levinson, A. D. (1983). Nucleotide sequence to the v-myc oncogene of avian retrovirus MC29. Proc Natl Acad Sci U S A 80, 100-104. Alves, N. L., Derks, I. A., Berk, E., Spijker, R., van Lier, R. A., and Eldering, E. (2006). The Noxa/Mcl-1 axis regulates susceptibility to apoptosis under glucose limitation in dividing T cells. Immunity 24, 703-716. Amati, B., Littlewood, T. D., Evan, G. I., and Land, H. (1993). The c-Myc protein induces cell cycle progression and apoptosis through dimerization with Max. Embo J 12, 5083-5087. Annis, M. G., Soucie, E. L., Dlugosz, P. J., Cruz-Aguado, J. A., Penn, L. Z., Leber, B., and Andrews, D. W. (2005). Bax forms multispanning monomers that oligomerize to permeabilize membranes during apoptosis. Embo J 24, 2096-2103. Arvanitis, C., and Felsher, D. W. (2006). Conditional transgenic models define how MYC initiates and maintains tumorigenesis. Semin Cancer Biol 16, 313-317. Ashkenazi, A., and Dixit, V. M. (1998). Death receptors: signaling and modulation. Science 281, 1305-1308. Askew, D. S., Ashmun, R. A., Simmons, B. C., and Cleveland, J. L. (1991). Constitutive c-myc expression in an IL-3-dependent myeloid cell line suppresses cell cycle arrest and accelerates apoptosis. Oncogene 6, 1915-1922. Baas, A. F., Kuipers, J., van der Wel, N. N., Batlle, E., Koerten, H. K., Peters, P. J., and Clevers, H. C. (2004a). Complete polarization of single intestinal epithelial cells upon activation of LKB1 by STRAD. Cell 116, 457-466. Baas, A. F., Smit, L., and Clevers, H. (2004b). LKB1 tumor suppressor protein: PARtaker in cell polarity. Trends in cell biology 14, 312-319. Bartek, J., Bartkova, J., Vojtesek, B., Staskova, Z., Lukas, J., Rejthar, A., Kovarik, J., Midgley, C. A., Gannon, J. V., and Lane, D. P. (1991). Aberrant expression of the p53 oncoprotein is a common feature of a wide spectrum of human malignancies. Oncogene 6, 1699-1703. Basu, A., and Haldar, S. (2003). Identification of a novel Bcl-xL phosphorylation site regulating the sensitivity of taxol- or 2-methoxyestradiol-induced apoptosis. FEBS Lett 538, 41-47. Benchimol, S. (2001). p53-dependent pathways of apoptosis. Cell Death Differ 8, 1049-1051. Bensaad, K., Tsuruta, A., Selak, M. A., Vidal, M. N., Nakano, K., Bartrons, R., Gottlieb, E., and Vousden, K. H. (2006). TIGAR, a p53-inducible regulator of glycolysis and apoptosis. Cell 126, 107-120. Bentley, D. L., and Groudine, M. (1986). A block to elongation is largely responsible for decreased transcription of c-myc in differentiated HL60 cells. Nature 321, 702-706. Bergsbaken, T., Fink, S. L., and Cookson, B. T. (2009). Pyroptosis: host cell death and inflammation. Nature reviews Microbiology 7, 99-109. Berkers, C. R., Maddocks, O. D., Cheung, E. C., Mor, I., and Vousden, K. H. (2013). Metabolic regulation by p53 family members. Cell Metab 18, 617-633. Bhatia, K., Huppi, K., Spangler, G., Siwarski, D., Iyer, R., and Magrath, I. (1993). Point mutations in the c-Myc transactivation domain are common in Burkitt's lymphoma and mouse plasmacytomas. Nat Genet 5, 56-61. Bieche, I., Laurendeau, I., Tozlu, S., Olivi, M., Vidaud, D., Lidereau, R., and Vidaud, M. (1999). Quantitation of MYC gene expression in sporadic breast tumors with a real-time reverse transcription-PCR assay. Cancer Res 59, 2759-2765. Blackwell, T. K., Kretzner, L., Blackwood, E. M., Eisenman, R. N., and Weintraub, H. (1990). Sequence-specific DNA binding by the c-Myc protein. Science 250, 1149-1151. Blackwood, E. M., and Eisenman, R. N. (1991). Max: a helix-loop-helix zipper protein that forms a sequence-specific DNA-binding complex with Myc. Science 251, 1211-1217. Bleicken, S., Classen, M., Padmavathi, P. V., Ishikawa, T., Zeth, K., Steinhoff, H. J., and Bordignon, E. (2010). Molecular details of Bax activation, oligomerization, and membrane insertion. J Biol Chem 285, 6636-6647. Bode, A. M., and Dong, Z. (2004). Post-translational modification of p53 in tumorigenesis. Nat Rev Cancer 4, 793-805.

Page 90: Apoptosis by c-Myc - Helsinki

Boldin, M. P., Varfolomeev, E. E., Pancer, Z., Mett, I. L., Camonis, J. H., and Wallach, D. (1995). A novel protein that interacts with the death domain of Fas/APO1 contains a sequence motif related to the death domain. J Biol Chem 270, 7795-7798. Bratton, S. B., and Salvesen, G. S. (2010). Regulation of the Apaf-1-caspase-9 apoptosome. Journal of cell science 123, 3209-3214. Broaddus, V. C., Dansen, T. B., Abayasiriwardana, K. S., Wilson, S. M., Finch, A. J., Swigart, L. B., Hunt, A. E., and Evan, G. I. (2005). Bid mediates apoptotic synergy between tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) and DNA damage. J Biol Chem 280, 12486-12493. Brosh, R., and Rotter, V. (2009). When mutants gain new powers: news from the mutant p53 field. Nat Rev Cancer 9, 701-713. Brown, C. J., Lain, S., Verma, C. S., Fersht, A. R., and Lane, D. P. (2009). Awakening guardian angels: drugging the p53 pathway. Nat Rev Cancer 9, 862-873. Buchakjian, M. R., and Kornbluth, S. (2010). The engine driving the ship: metabolic steering of cell proliferation and death. Nature reviews Molecular cell biology 11, 715-727. Cairns, R. A., Harris, I. S., and Mak, T. W. (2011). Regulation of cancer cell metabolism. Nat Rev Cancer 11, 85-95. Chan, F. K., Chun, H. J., Zheng, L., Siegel, R. M., Bui, K. L., and Lenardo, M. J. (2000). A domain in TNF receptors that mediates ligand-independent receptor assembly and signaling. Science 288, 2351-2354. Chang, T. C., Yu, D., Lee, Y. S., Wentzel, E. A., Arking, D. E., West, K. M., Dang, C. V., Thomas-Tikhonenko, A., and Mendell, J. T. (2008). Widespread microRNA repression by Myc contributes to tumorigenesis. Nat Genet 40, 43-50. Chen, L., Willis, S. N., Wei, A., Smith, B. J., Fletcher, J. I., Hinds, M. G., Colman, P. M., Day, C. L., Adams, J. M., and Huang, D. C. (2005). Differential targeting of prosurvival Bcl-2 proteins by their BH3-only ligands allows complementary apoptotic function. Mol Cell 17, 393-403. Cheng, S. W., Davies, K. P., Yung, E., Beltran, R. J., Yu, J., and Kalpana, G. V. (1999). c-MYC interacts with INI1/hSNF5 and requires the SWI/SNF complex for transactivation function. Nat Genet 22, 102-105. Chinnaiyan, A. M., O'Rourke, K., Tewari, M., and Dixit, V. M. (1995). FADD, a novel death domain-containing protein, interacts with the death domain of Fas and initiates apoptosis. Cell 81, 505-512. Chipuk, J. E., Bouchier-Hayes, L., Kuwana, T., Newmeyer, D. D., and Green, D. R. (2005). PUMA couples the nuclear and cytoplasmic proapoptotic function of p53. Science 309, 1732-1735. Chipuk, J. E., Kuwana, T., Bouchier-Hayes, L., Droin, N. M., Newmeyer, D. D., Schuler, M., and Green, D. R. (2004). Direct activation of Bax by p53 mediates mitochondrial membrane permeabilization and apoptosis. Science 303, 1010-1014. Chipuk, J. E., Moldoveanu, T., Llambi, F., Parsons, M. J., and Green, D. R. (2010). The BCL-2 family reunion. Mol Cell 37, 299-310. Christofori, G., and Semb, H. (1999). The role of the cell-adhesion molecule E-cadherin as a tumour-suppressor gene. Trends in biochemical sciences 24, 73-76. Chrzan, P., Skokowski, J., Karmolinski, A., and Pawelczyk, T. (2001). Amplification of c-myc gene and overexpression of c-Myc protein in breast cancer and adjacent non-neoplastic tissue. Clin Biochem 34, 557-562. Conde, E., Suarez-Gauthier, A., Garcia-Garcia, E., Lopez-Rios, F., Lopez-Encuentra, A., Garcia-Lujan, R., Morente, M., Sanchez-Verde, L., and Sanchez-Cespedes, M. (2007). Specific pattern of LKB1 and phospho-acetyl-CoA carboxylase protein immunostaining in human normal tissues and lung carcinomas. Hum Pathol 38, 1351-1360. Counter, C. M., Hahn, W. C., Wei, W., Caddle, S. D., Beijersbergen, R. L., Lansdorp, P. M., Sedivy, J. M., and Weinberg, R. A. (1998). Dissociation among in vitro telomerase activity, telomere maintenance, and cellular immortalization. Proc Natl Acad Sci U S A 95, 14723-14728. Dang, C. V. (2012). MYC on the path to cancer. Cell 149, 22-35. Dang, C. V. (2013). MYC, metabolism, cell growth, and tumorigenesis. Cold Spring Harbor perspectives in medicine 3. Dang, C. V., McGuire, M., Buckmire, M., and Lee, W. M. (1989). Involvement of the 'leucine zipper' region in the oligomerization and transforming activity of human c-myc protein. Nature 337, 664-666. Dang, C. V., O'Donnell K, A., and Juopperi, T. (2005). The great MYC escape in tumorigenesis. Cancer Cell 8, 177-178. Dani, C., Blanchard, J. M., Piechaczyk, M., El Sabouty, S., Marty, L., and Jeanteur, P. (1984). Extreme instability of myc mRNA in normal and transformed human cells. Proc Natl Acad Sci U S A 81, 7046-7050. Danial, N. N. (2007). BCL-2 family proteins: critical checkpoints of apoptotic cell death. Clin Cancer Res 13, 7254-7263.

Page 91: Apoptosis by c-Myc - Helsinki

Danial, N. N., Gramm, C. F., Scorrano, L., Zhang, C. Y., Krauss, S., Ranger, A. M., Datta, S. R., Greenberg, M. E., Licklider, L. J., Lowell, B. B., et al. (2003). BAD and glucokinase reside in a mitochondrial complex that integrates glycolysis and apoptosis. Nature 424, 952-956. Dansen, T. B., Whitfield, J., Rostker, F., Brown-Swigart, L., and Evan, G. I. (2006). Specific requirement for Bax, not Bak, in Myc-induced apoptosis and tumor suppression in vivo. J Biol Chem 281, 10890-10895. David, C. J., Chen, M., Assanah, M., Canoll, P., and Manley, J. L. (2010). HnRNP proteins controlled by c-Myc deregulate pyruvate kinase mRNA splicing in cancer. Nature 463, 364-368. Davis, A. C., Wims, M., Spotts, G. D., Hann, S. R., and Bradley, A. (1993). A null c-myc mutation causes lethality before 10.5 days of gestation in homozygotes and reduced fertility in heterozygous female mice. Genes Dev 7, 671-682. de Klein, A., van Kessel, A. G., Grosveld, G., Bartram, C. R., Hagemeijer, A., Bootsma, D., Spurr, N. K., Heisterkamp, N., Groffen, J., and Stephenson, J. R. (1982). A cellular oncogene is translocated to the Philadelphia chromosome in chronic myelocytic leukaemia. Nature 300, 765-767. DeBerardinis, R. J., Lum, J. J., Hatzivassiliou, G., and Thompson, C. B. (2008). The biology of cancer: metabolic reprogramming fuels cell growth and proliferation. Cell Metab 7, 11-20. Degterev, A., Boyce, M., and Yuan, J. (2003). A decade of caspases. Oncogene 22, 8543-8567. Degterev, A., and Yuan, J. (2008). Expansion and evolution of cell death programmes. Nature reviews Molecular cell biology 9, 378-390. Denault, J. B., and Salvesen, G. S. (2002). Caspases: keys in the ignition of cell death. Chem Rev 102, 4489-4500. Deveraux, Q. L., Roy, N., Stennicke, H. R., Van Arsdale, T., Zhou, Q., Srinivasula, S. M., Alnemri, E. S., Salvesen, G. S., and Reed, J. C. (1998). IAPs block apoptotic events induced by caspase-8 and cytochrome c by direct inhibition of distinct caspases. Embo J 17, 2215-2223. Deveraux, Q. L., Takahashi, R., Salvesen, G. S., and Reed, J. C. (1997). X-linked IAP is a direct inhibitor of cell-death proteases. Nature 388, 300-304. Dews, M., Fox, J. L., Hultine, S., Sundaram, P., Wang, W., Liu, Y. Y., Furth, E., Enders, G. H., El-Deiry, W., Schelter, J. M., et al. (2010). The myc-miR-17~92 axis blunts TGF{beta} signaling and production of multiple TGF{beta}-dependent antiangiogenic factors. Cancer Res 70, 8233-8246. Dewson, G., Kratina, T., Czabotar, P., Day, C. L., Adams, J. M., and Kluck, R. M. (2009). Bak activation for apoptosis involves oligomerization of dimers via their alpha6 helices. Mol Cell 36, 696-703. Dewson, G., Kratina, T., Sim, H. W., Puthalakath, H., Adams, J. M., Colman, P. M., and Kluck, R. M. (2008). To trigger apoptosis, Bak exposes its BH3 domain and homodimerizes via BH3:groove interactions. Mol Cell 30, 369-380. Ding, J., Zhang, Z., Roberts, G. J., Falcone, M., Miao, Y., Shao, Y., Zhang, X. C., Andrews, D. W., and Lin, J. (2010). Bcl-2 and Bax interact via the BH1-3 groove-BH3 motif interface and a novel interface involving the BH4 motif. J Biol Chem 285, 28749-28763. Drubin, D. G., and Nelson, W. J. (1996). Origins of cell polarity. Cell 84, 335-344. Egle, A., Harris, A. W., Bouillet, P., and Cory, S. (2004). Bim is a suppressor of Myc-induced mouse B cell leukemia. Proc Natl Acad Sci U S A 101, 6164-6169. Eilers, M., Picard, D., Yamamoto, K. R., and Bishop, J. M. (1989). Chimaeras of myc oncoprotein and steroid receptors cause hormone-dependent transformation of cells. Nature 340, 66-68. Eilers, M., Schirm, S., and Bishop, J. M. (1991). The MYC protein activates transcription of the alpha-prothymosin gene. EMBO J 10, 133-141. Eischen, C. M., Roussel, M. F., Korsmeyer, S. J., and Cleveland, J. L. (2001a). Bax loss impairs Myc-induced apoptosis and circumvents the selection of p53 mutations during Myc-mediated lymphomagenesis. Mol Cell Biol 21, 7653-7662. Eischen, C. M., Weber, J. D., Roussel, M. F., Sherr, C. J., and Cleveland, J. L. (1999). Disruption of the ARF-Mdm2-p53 tumor suppressor pathway in Myc-induced lymphomagenesis. Genes Dev 13, 2658-2669. Eischen, C. M., Woo, D., Roussel, M. F., and Cleveland, J. L. (2001b). Apoptosis triggered by Myc-induced suppression of Bcl-X(L) or Bcl-2 is bypassed during lymphomagenesis. Mol Cell Biol 21, 5063-5070. Ellis, H. M., and Horvitz, H. R. (1986). Genetic control of programmed cell death in the nematode C. elegans. Cell 44, 817-829. Ellwood-Yen, K., Graeber, T. G., Wongvipat, J., Iruela-Arispe, M. L., Zhang, J., Matusik, R., Thomas, G. V., and Sawyers, C. L. (2003). Myc-driven murine prostate cancer shares molecular features with human prostate tumors. Cancer Cell 4, 223-238. Elson, A., Deng, C., Campos-Torres, J., Donehower, L. A., and Leder, P. (1995). The MMTV/c-myc transgene and p53 null alleles collaborate to induce T-cell lymphomas, but not mammary carcinomas in transgenic mice. Oncogene 11, 181-190.

Page 92: Apoptosis by c-Myc - Helsinki

Emery, J. G., McDonnell, P., Burke, M. B., Deen, K. C., Lyn, S., Silverman, C., Dul, E., Appelbaum, E. R., Eichman, C., DiPrinzio, R., et al. (1998). Osteoprotegerin is a receptor for the cytotoxic ligand TRAIL. J Biol Chem 273, 14363-14367. Enari, M., Sakahira, H., Yokoyama, H., Okawa, K., Iwamatsu, A., and Nagata, S. (1998). A caspase-activated DNase that degrades DNA during apoptosis, and its inhibitor ICAD. Nature 391, 43-50. Erisman, M. D., Rothberg, P. G., Diehl, R. E., Morse, C. C., Spandorfer, J. M., and Astrin, S. M. (1985). Deregulation of c-myc gene expression in human colon carcinoma is not accompanied by amplification or rearrangement of the gene. Mol Cell Biol 5, 1969-1976. Erster, S., Mihara, M., Kim, R. H., Petrenko, O., and Moll, U. M. (2004). In vivo mitochondrial p53 translocation triggers a rapid first wave of cell death in response to DNA damage that can precede p53 target gene activation. Mol Cell Biol 24, 6728-6741. Evan, G. I., Wyllie, A. H., Gilbert, C. S., Littlewood, T. D., Land, H., Brooks, M., Waters, C. M., Penn, L. Z., and Hancock, D. C. (1992). Induction of apoptosis in fibroblasts by c-myc protein. Cell 69, 119-128. Fanidi, A., Harrington, E. A., and Evan, G. I. (1992). Cooperative interaction between c-myc and bcl-2 proto-oncogenes. Nature 359, 554-556. Felsher, D. W., and Bishop, J. M. (1999). Transient excess of MYC activity can elicit genomic instability and tumorigenesis. Proc Natl Acad Sci U S A 96, 3940-3944. Feng, L., Hollstein, M., and Xu, Y. (2006). Ser46 phosphorylation regulates p53-dependent apoptosis and replicative senescence. Cell Cycle 5, 2812-2819. Fink, S. L., and Cookson, B. T. (2005). Apoptosis, pyroptosis, and necrosis: mechanistic description of dead and dying eukaryotic cells. Infection and immunity 73, 1907-1916. Fischer, U., Janicke, R. U., and Schulze-Osthoff, K. (2003). Many cuts to ruin: a comprehensive update of caspase substrates. Cell Death Differ 10, 76-100. Foley, K. P., and Eisenman, R. N. (1999). Two MAD tails: what the recent knockouts of Mad1 and Mxi1 tell us about the MYC/MAX/MAD network. Biochim Biophys Acta 1423, M37-47. Fridman, J. S., and Lowe, S. W. (2003). Control of apoptosis by p53. Oncogene 22, 9030-9040. Fulda, S., and Debatin, K. M. (2006). Extrinsic versus intrinsic apoptosis pathways in anticancer chemotherapy. Oncogene 25, 4798-4811. Fulda, S., Meyer, E., and Debatin, K. M. (2002). Inhibition of TRAIL-induced apoptosis by Bcl-2 overexpression. Oncogene 21, 2283-2294. Galluzzi, L., Kepp, O., Heiden, M. G., and Kroemer, G. (2013). Metabolic targets for cancer therapy. Nature reviews Drug discovery 12, 829-846. Galluzzi, L., and Kroemer, G. (2008). Necroptosis: a specialized pathway of programmed necrosis. Cell 135, 1161-1163. Gandarillas, A., and Watt, F. M. (1997). c-Myc promotes differentiation of human epidermal stem cells. Genes Dev 11, 2869-2882. Gao, P., Tchernyshyov, I., Chang, T. C., Lee, Y. S., Kita, K., Ochi, T., Zeller, K. I., De Marzo, A. M., Van Eyk, J. E., Mendell, J. T., and Dang, C. V. (2009). c-Myc suppression of miR-23a/b enhances mitochondrial glutaminase expression and glutamine metabolism. Nature 458, 762-765. Gatenby, R. A., and Gillies, R. J. (2004). Why do cancers have high aerobic glycolysis? Nat Rev Cancer 4, 891-899. Gavathiotis, E., Suzuki, M., Davis, M. L., Pitter, K., Bird, G. H., Katz, S. G., Tu, H. C., Kim, H., Cheng, E. H., Tjandra, N., and Walensky, L. D. (2008). BAX activation is initiated at a novel interaction site. Nature 455, 1076-1081. Golstein, P., and Kroemer, G. (2007). Cell death by necrosis: towards a molecular definition. Trends in biochemical sciences 32, 37-43. Gonzalez-Angulo, A. M., Morales-Vasquez, F., and Hortobagyi, G. N. (2007). Overview of resistance to systemic therapy in patients with breast cancer. Advances in experimental medicine and biology 608, 1-22. Graves, J. A., Wang, Y., Sims-Lucas, S., Cherok, E., Rothermund, K., Branca, M. F., Elster, J., Beer-Stolz, D., Van Houten, B., Vockley, J., and Prochownik, E. V. (2012). Mitochondrial structure, function and dynamics are temporally controlled by c-Myc. PLoS One 7, e37699. Hagn, F., Klein, C., Demmer, O., Marchenko, N., Vaseva, A., Moll, U. M., and Kessler, H. (2010). BclxL changes conformation upon binding to wild-type but not mutant p53 DNA binding domain. J Biol Chem 285, 3439-3450. Hanahan, D., and Weinberg, R. A. (2000). The hallmarks of cancer. Cell 100, 57-70. Hanahan, D., and Weinberg, R. A. (2011). Hallmarks of cancer: the next generation. Cell 144, 646-674. Hann, S. R. (2006). Role of post-translational modifications in regulating c-Myc proteolysis, transcriptional activity and biological function. Semin Cancer Biol 16, 288-302.

Page 93: Apoptosis by c-Myc - Helsinki

Hardie, D. G. (2007). AMP-activated/SNF1 protein kinases: conserved guardians of cellular energy. Nature reviews Molecular cell biology 8, 774-785. Hardie, D. G. (2011). Sensing of energy and nutrients by AMP-activated protein kinase. Am J Clin Nutr 93, 891S-896. Hardie, D. G., Ross, F. A., and Hawley, S. A. (2012). AMPK: a nutrient and energy sensor that maintains energy homeostasis. Nature reviews Molecular cell biology 13, 251-262. Harrington, E. A., Bennett, M. R., Fanidi, A., and Evan, G. I. (1994). c-Myc-induced apoptosis in fibroblasts is inhibited by specific cytokines. Embo J 13, 3286-3295. Harris, A. L. (2002). Hypoxia--a key regulatory factor in tumour growth. Nat Rev Cancer 2, 38-47. Hawley, S. A., Davison, M., Woods, A., Davies, S. P., Beri, R. K., Carling, D., and Hardie, D. G. (1996). Characterization of the AMP-activated protein kinase kinase from rat liver and identification of threonine 172 as the major site at which it phosphorylates AMP-activated protein kinase. J Biol Chem 271, 27879-27887. Hawley, S. A., Pan, D. A., Mustard, K. J., Ross, L., Bain, J., Edelman, A. M., Frenguelli, B. G., and Hardie, D. G. (2005). Calmodulin-dependent protein kinase kinase-beta is an alternative upstream kinase for AMP-activated protein kinase. Cell Metab 2, 9-19. Hay, B. A., and Guo, M. (2006). Caspase-dependent cell death in Drosophila. Annual review of cell and developmental biology 22, 623-650. Hearle, N., Schumacher, V., Menko, F. H., Olschwang, S., Boardman, L. A., Gille, J. J., Keller, J. J., Westerman, A. M., Scott, R. J., Lim, W., et al. (2006). Frequency and spectrum of cancers in the Peutz-Jeghers syndrome. Clin Cancer Res 12, 3209-3215. Hemann, M. T., Bric, A., Teruya-Feldstein, J., Herbst, A., Nilsson, J. A., Cordon-Cardo, C., Cleveland, J. L., Tansey, W. P., and Lowe, S. W. (2005). Evasion of the p53 tumour surveillance network by tumour-derived MYC mutants. Nature 436, 807-811. Hermeking, H., and Eick, D. (1994). Mediation of c-Myc-induced apoptosis by p53. Science 265, 2091-2093. Hsu, B., Marin, M. C., el-Naggar, A. K., Stephens, L. C., Brisbay, S., and McDonnell, T. J. (1995). Evidence that c-myc mediated apoptosis does not require wild-type p53 during lymphomagenesis. Oncogene 11, 175-179. Hsu, Y. T., Wolter, K. G., and Youle, R. J. (1997). Cytosol-to-membrane redistribution of Bax and Bcl-X(L) during apoptosis. Proc Natl Acad Sci U S A 94, 3668-3672. Huang, D. C., and Strasser, A. (2000). BH3-Only proteins-essential initiators of apoptotic cell death. Cell 103, 839-842. Hueber, A. O., Zornig, M., Lyon, D., Suda, T., Nagata, S., and Evan, G. I. (1997). Requirement for the CD95 receptor-ligand pathway in c-Myc-induced apoptosis. Science 278, 1305-1309. Humbert, N., Navaratnam, N., Augert, A., Da Costa, M., Martien, S., Wang, J., Martinez, D., Abbadie, C., Carling, D., de Launoit, Y., et al. (2010). Regulation of ploidy and senescence by the AMPK-related kinase NUAK1. EMBO J 29, 376-386. Hunter, T. (1997). Oncoprotein networks. Cell 88, 333-346. Imamura, K., Ogura, T., Kishimoto, A., Kaminishi, M., and Esumi, H. (2001). Cell cycle regulation via p53 phosphorylation by a 5'-AMP activated protein kinase activator, 5-aminoimidazole- 4-carboxamide-1-beta-D-ribofuranoside, in a human hepatocellular carcinoma cell line. Biochem Biophys Res Commun 287, 562-567. Inoki, K., Zhu, T., and Guan, K. L. (2003). TSC2 mediates cellular energy response to control cell growth and survival. Cell 115, 577-590. Iritani, B. M., and Eisenman, R. N. (1999). c-Myc enhances protein synthesis and cell size during B lymphocyte development. Proc Natl Acad Sci U S A 96, 13180-13185. Janicke, R. U., Lin, X. Y., Lee, F. H., and Porter, A. G. (1996). Cyclin D3 sensitizes tumor cells to tumor necrosis factor-induced, c-Myc-dependent apoptosis. Mol Cell Biol 16, 5245-5253. Jeghers, H., Mc, K. V., and Katz, K. H. (1949). Generalized intestinal polyposis and melanin spots of the oral mucosa, lips and digits; a syndrome of diagnostic significance. N Engl J Med 241, 993, illust; passim. Jiang, W., Zhu, Z., and Thompson, H. J. (2008). Dietary energy restriction modulates the activity of AMP-activated protein kinase, Akt, and mammalian target of rapamycin in mammary carcinomas, mammary gland, and liver. Cancer Res 68, 5492-5499. Johnston, L. A., Prober, D. A., Edgar, B. A., Eisenman, R. N., and Gallant, P. (1999). Drosophila myc regulates cellular growth during development. Cell 98, 779-790. Jones, R. G., Plas, D. R., Kubek, S., Buzzai, M., Mu, J., Xu, Y., Birnbaum, M. J., and Thompson, C. B. (2005). AMP-activated protein kinase induces a p53-dependent metabolic checkpoint. Mol Cell 18, 283-293.

Page 94: Apoptosis by c-Myc - Helsinki

Juin, P., Hueber, A. O., Littlewood, T., and Evan, G. (1999). c-Myc-induced sensitization to apoptosis is mediated through cytochrome c release. Genes Dev 13, 1367-1381. Junttila, M. R., and Evan, G. I. (2009). p53--a Jack of all trades but master of none. Nat Rev Cancer 9, 821-829. Kandasamy, K., Srinivasula, S. M., Alnemri, E. S., Thompson, C. B., Korsmeyer, S. J., Bryant, J. L., and Srivastava, R. K. (2003). Involvement of proapoptotic molecules Bax and Bak in tumor necrosis factor-related apoptosis-inducing ligand (TRAIL)-induced mitochondrial disruption and apoptosis: differential regulation of cytochrome c and Smac/DIABLO release. Cancer Res 63, 1712-1721. Karbowski, M., Norris, K. L., Cleland, M. M., Jeong, S. Y., and Youle, R. J. (2006). Role of Bax and Bak in mitochondrial morphogenesis. Nature 443, 658-662. Karlsson, A., Giuriato, S., Tang, F., Fung-Weier, J., Levan, G., and Felsher, D. W. (2003). Genomically complex lymphomas undergo sustained tumor regression upon MYC inactivation unless they acquire novel chromosomal translocations. Blood 101, 2797-2803. Kastan, M. B., Onyekwere, O., Sidransky, D., Vogelstein, B., and Craig, R. W. (1991). Participation of p53 protein in the cellular response to DNA damage. Cancer Res 51, 6304-6311. Kauffmann-Zeh, A., Rodriguez-Viciana, P., Ulrich, E., Gilbert, C., Coffer, P., Downward, J., and Evan, G. (1997). Suppression of c-Myc-induced apoptosis by Ras signalling through PI(3)K and PKB. Nature 385, 544-548. Kelly, P. N., Grabow, S., Delbridge, A. R., Strasser, A., and Adams, J. M. (2011). Endogenous Bcl-xL is essential for Myc-driven lymphomagenesis in mice. Blood. Kerr, J. F., Wyllie, A. H., and Currie, A. R. (1972). Apoptosis: a basic biological phenomenon with wide-ranging implications in tissue kinetics. Br J Cancer 26, 239-257. Kim, H., Tu, H. C., Ren, D., Takeuchi, O., Jeffers, J. R., Zambetti, G. P., Hsieh, J. J., and Cheng, E. H. (2009). Stepwise activation of BAX and BAK by tBID, BIM, and PUMA initiates mitochondrial apoptosis. Mol Cell 36, 487-499. Kim, J. W., and Dang, C. V. (2006). Cancer's molecular sweet tooth and the Warburg effect. Cancer Res 66, 8927-8930. Kischkel, F. C., Lawrence, D. A., Chuntharapai, A., Schow, P., Kim, K. J., and Ashkenazi, A. (2000). Apo2L/TRAIL-dependent recruitment of endogenous FADD and caspase-8 to death receptors 4 and 5. Immunity 12, 611-620. Klefstrom, J., Arighi, E., Littlewood, T., Jaattela, M., Saksela, E., Evan, G. I., and Alitalo, K. (1997). Induction of TNF-sensitive cellular phenotype by c-Myc involves p53 and impaired NF-kappaB activation. Embo J 16, 7382-7392. Klefstrom, J., Vastrik, I., Saksela, E., Valle, J., Eilers, M., and Alitalo, K. (1994). c-Myc induces cellular susceptibility to the cytotoxic action of TNF- alpha. Embo J 13, 5442-5450. Klefstrom, J., Verschuren, E. W., and Evan, G. (2002). c-Myc Augments the Apoptotic Activity of Cytosolic Death Receptor Signaling Proteins by Engaging the Mitochondrial Apoptotic Pathway. J Biol Chem 277, 43224-43232. Kleine-Kohlbrecher, D., Adhikary, S., and Eilers, M. (2006). Mechanisms of transcriptional repression by Myc. Current topics in microbiology and immunology 302, 51-62. Knudson, A. G., Jr. (1971). Mutation and cancer: statistical study of retinoblastoma. Proc Natl Acad Sci U S A 68, 820-823. Konopleva, M., Contractor, R., Tsao, T., Samudio, I., Ruvolo, P. P., Kitada, S., Deng, X., Zhai, D., Shi, Y. X., Sneed, T., et al. (2006). Mechanisms of apoptosis sensitivity and resistance to the BH3 mimetic ABT-737 in acute myeloid leukemia. Cancer Cell 10, 375-388. Kress, M., May, E., Cassingena, R., and May, P. (1979). Simian virus 40-transformed cells express new species of proteins precipitable by anti-simian virus 40 tumor serum. Journal of virology 31, 472-483. Kroemer, G., Galluzzi, L., and Brenner, C. (2007). Mitochondrial membrane permeabilization in cell death. Physiological reviews 87, 99-163. Kroemer, G., and Pouyssegur, J. (2008). Tumor cell metabolism: cancer's Achilles' heel. Cancer Cell 13, 472-482. Kuwana, T., Bouchier-Hayes, L., Chipuk, J. E., Bonzon, C., Sullivan, B. A., Green, D. R., and Newmeyer, D. D. (2005). BH3 domains of BH3-only proteins differentially regulate Bax-mediated mitochondrial membrane permeabilization both directly and indirectly. Mol Cell 17, 525-535. Kuwana, T., Mackey, M. R., Perkins, G., Ellisman, M. H., Latterich, M., Schneiter, R., Green, D. R., and Newmeyer, D. D. (2002). Bid, Bax, and lipids cooperate to form supramolecular openings in the outer mitochondrial membrane. Cell 111, 331-342. Land, H., Parada, L. F., and Weinberg, R. A. (1983). Tumorigenic conversion of primary embryo fibroblasts requires at least two cooperating oncogenes. Nature 304, 596-602.

Page 95: Apoptosis by c-Myc - Helsinki

Lane, D. P. (1992). Cancer. p53, guardian of the genome. Nature 358, 15-16. Lane, D. P., and Benchimol, S. (1990). p53: oncogene or anti-oncogene? Genes Dev 4, 1-8. Lane, D. P., and Crawford, L. V. (1979). T antigen is bound to a host protein in SV40-transformed cells. Nature 278, 261-263. Langdon, W. Y., Harris, A. W., Cory, S., and Adams, J. M. (1986). The c-myc oncogene perturbs B lymphocyte development in E-mu-myc transgenic mice. Cell 47, 11-18. Leber, B., Lin, J., and Andrews, D. W. (2007). Embedded together: the life and death consequences of interaction of the Bcl-2 family with membranes. Apoptosis 12, 897-911. LeBlanc, H., Lawrence, D., Varfolomeev, E., Totpal, K., Morlan, J., Schow, P., Fong, S., Schwall, R., Sinicropi, D., and Ashkenazi, A. (2002). Tumor-cell resistance to death receptor--induced apoptosis through mutational inactivation of the proapoptotic Bcl-2 homolog Bax. Nat Med 8, 274-281. Leder, A., Pattengale, P. K., Kuo, A., Stewart, T. A., and Leder, P. (1986). Consequences of widespread deregulation of the c-myc gene in transgenic mice: multiple neoplasms and normal development. Cell 45, 485-495. Lee, J. H., Koh, H., Kim, M., Kim, Y., Lee, S. Y., Karess, R. E., Lee, S. H., Shong, M., Kim, J. M., Kim, J., and Chung, J. (2007). Energy-dependent regulation of cell structure by AMP-activated protein kinase. Nature 447, 1017-1020. Lee, T., Yao, G., Nevins, J., and You, L. (2008). Sensing and integration of Erk and PI3K signals by Myc. PLoS computational biology 4, e1000013. Letai, A., Bassik, M. C., Walensky, L. D., Sorcinelli, M. D., Weiler, S., and Korsmeyer, S. J. (2002). Distinct BH3 domains either sensitize or activate mitochondrial apoptosis, serving as prototype cancer therapeutics. Cancer Cell 2, 183-192. Levine, A. J. (1997). p53, the cellular gatekeeper for growth and division. Cell 88, 323-331. Li, B., and Simon, M. C. (2013). Molecular Pathways: Targeting MYC-induced Metabolic Reprogramming and Oncogenic Stress in Cancer. Clin Cancer Res 19, 5835-5841. Li, F., Wang, Y., Zeller, K. I., Potter, J. J., Wonsey, D. R., O'Donnell, K. A., Kim, J. W., Yustein, J. T., Lee, L. A., and Dang, C. V. (2005). Myc stimulates nuclearly encoded mitochondrial genes and mitochondrial biogenesis. Mol Cell Biol 25, 6225-6234. Liang, J., and Mills, G. B. (2013). AMPK: a contextual oncogene or tumor suppressor? Cancer Res 73, 2929-2935. Lindstrom, M. S., and Wiman, K. G. (2003). Myc and E2F1 induce p53 through p14ARF-independent mechanisms in human fibroblasts. Oncogene 22, 4993-5005. Littlewood, T. D., Hancock, D. C., Danielian, P. S., Parker, M. G., and Evan, G. I. (1995). A modified oestrogen receptor ligand-binding domain as an improved switch for the regulation of heterologous proteins. Nucleic acids research 23, 1686-1690. Liu, L., Ulbrich, J., Muller, J., Wustefeld, T., Aeberhard, L., Kress, T. R., Muthalagu, N., Rycak, L., Rudalska, R., Moll, R., et al. (2012). Deregulated MYC expression induces dependence upon AMPK-related kinase 5. Nature 483, 608-612. Loeb, L. A., Loeb, K. R., and Anderson, J. P. (2003). Multiple mutations and cancer. Proc Natl Acad Sci U S A 100, 776-781. Lomonosova, E., and Chinnadurai, G. (2008). BH3-only proteins in apoptosis and beyond: an overview. Oncogene 27 Suppl 1, S2-19. Lowe, S. W., Cepero, E., and Evan, G. (2004). Intrinsic tumour suppression. Nature 432, 307-315. Lowe, S. W., and Sherr, C. J. (2003). Tumor suppression by Ink4a-Arf: progress and puzzles. Current opinion in genetics & development 13, 77-83. Luo, X., Budihardjo, I., Zou, H., Slaughter, C., and Wang, X. (1998). Bid, a Bcl2 interacting protein, mediates cytochrome c release from mitochondria in response to activation of cell surface death receptors. Cell 94, 481-490. Luo, Z., Zang, M., and Guo, W. (2010). AMPK as a metabolic tumor suppressor: control of metabolism and cell growth. Future Oncol 6, 457-470. Lutz, W., Leon, J., and Eilers, M. (2002). Contributions of Myc to tumorigenesis. Biochim Biophys Acta 1602, 61-71. MacLaine, N. J., and Hupp, T. R. (2011). How phosphorylation controls p53. Cell Cycle 10, 916-921. Maclean, K. H., Keller, U. B., Rodriguez-Galindo, C., Nilsson, J. A., and Cleveland, J. L. (2003). c-Myc augments gamma irradiation-induced apoptosis by suppressing Bcl-XL. Mol Cell Biol 23, 7256-7270. Maiuri, M. C., Zalckvar, E., Kimchi, A., and Kroemer, G. (2007). Self-eating and self-killing: crosstalk between autophagy and apoptosis. Nature reviews Molecular cell biology 8, 741-752. Malumbres, M., and Barbacid, M. (2003). RAS oncogenes: the first 30 years. Nat Rev Cancer 3, 459-465.

Page 96: Apoptosis by c-Myc - Helsinki

Mankoff, D. A., Eary, J. F., Link, J. M., Muzi, M., Rajendran, J. G., Spence, A. M., and Krohn, K. A. (2007). Tumor-specific positron emission tomography imaging in patients: [18F] fluorodeoxyglucose and beyond. Clin Cancer Res 13, 3460-3469. Mannava, S., Grachtchouk, V., Wheeler, L. J., Im, M., Zhuang, D., Slavina, E. G., Mathews, C. K., Shewach, D. S., and Nikiforov, M. A. (2008). Direct role of nucleotide metabolism in C-MYC-dependent proliferation of melanoma cells. Cell Cycle 7, 2392-2400. Marchenko, N. D., and Moll, U. M. (2007). The role of ubiquitination in the direct mitochondrial death program of p53. Cell Cycle 6, 1718-1723. Marchenko, N. D., Wolff, S., Erster, S., Becker, K., and Moll, U. M. (2007). Monoubiquitylation promotes mitochondrial p53 translocation. EMBO J 26, 923-934. Marsin, A. S., Bouzin, C., Bertrand, L., and Hue, L. (2002). The stimulation of glycolysis by hypoxia in activated monocytes is mediated by AMP-activated protein kinase and inducible 6-phosphofructo-2-kinase. J Biol Chem 277, 30778-30783. Martin, S. G., and St Johnston, D. (2003). A role for Drosophila LKB1 in anterior-posterior axis formation and epithelial polarity. Nature 421, 379-384. McCarthy, A., Lord, C. J., Savage, K., Grigoriadis, A., Smith, D. P., Weigelt, B., Reis-Filho, J. S., and Ashworth, A. (2009). Conditional deletion of the Lkb1 gene in the mouse mammary gland induces tumour formation. The Journal of pathology 219, 306-316. McGill, G., Shimamura, A., Bates, R. C., Savage, R. E., and Fisher, D. E. (1997). Loss of matrix adhesion triggers rapid transformation-selective apoptosis in fibroblasts. J Cell Biol 138, 901-911. McMahon, S. B., Wood, M. A., and Cole, M. D. (2000). The essential cofactor TRRAP recruits the histone acetyltransferase hGCN5 to c-Myc. Mol Cell Biol 20, 556-562. Medema, R. H., de Vries-Smits, A. M., van der Zon, G. C., Maassen, J. A., and Bos, J. L. (1993). Ras activation by insulin and epidermal growth factor through enhanced exchange of guanine nucleotides on p21ras. Mol Cell Biol 13, 155-162. Meyer, N., and Penn, L. Z. (2008). Reflecting on 25 years with MYC. Nat Rev Cancer 8, 976-990. Micheau, O., Thome, M., Schneider, P., Holler, N., Tschopp, J., Nicholson, D. W., Briand, C., and Grutter, M. G. (2002). The long form of FLIP is an activator of caspase-8 at the Fas death-inducing signaling complex. J Biol Chem 277, 45162-45171. Micheau, O., and Tschopp, J. (2003). Induction of TNF receptor I-mediated apoptosis via two sequential signaling complexes. Cell 114, 181-190. Mihara, M., Erster, S., Zaika, A., Petrenko, O., Chittenden, T., Pancoska, P., and Moll, U. M. (2003). p53 has a direct apoptogenic role at the mitochondria. Mol Cell 11, 577-590. Mikhailov, V., Mikhailova, M., Degenhardt, K., Venkatachalam, M. A., White, E., and Saikumar, P. (2003). Association of Bax and Bak homo-oligomers in mitochondria. Bax requirement for Bak reorganization and cytochrome c release. J Biol Chem 278, 5367-5376. Mirouse, V., and Billaud, M. (2011). The LKB1/AMPK polarity pathway. FEBS Lett 585, 981-985. Mitchell, K. O., Ricci, M. S., Miyashita, T., Dicker, D. T., Jin, Z., Reed, J. C., and El-Deiry, W. S. (2000). Bax is a transcriptional target and mediator of c-myc-induced apoptosis. Cancer Res 60, 6318-6325. Moldoveanu, T., Grace, C. R., Llambi, F., Nourse, A., Fitzgerald, P., Gehring, K., Kriwacki, R. W., and Green, D. R. (2013). BID-induced structural changes in BAK promote apoptosis. Nature structural & molecular biology 20, 589-597. Moll, U. M., Riou, G., and Levine, A. J. (1992). Two distinct mechanisms alter p53 in breast cancer: mutation and nuclear exclusion. Proc Natl Acad Sci U S A 89, 7262-7266. Mondal, S., Bhattacharya, K., Mallick, A., Sangwan, R., and Mandal, C. (2012). Bak compensated for Bax in p53-null cells to release cytochrome c for the initiation of mitochondrial signaling during Withanolide D-induced apoptosis. PLoS One 7, e34277. Morrish, F., Noonan, J., Perez-Olsen, C., Gafken, P. R., Fitzgibbon, M., Kelleher, J., VanGilst, M., and Hockenbery, D. (2010). Myc-dependent mitochondrial generation of acetyl-CoA contributes to fatty acid biosynthesis and histone acetylation during cell cycle entry. J Biol Chem 285, 36267-36274. Muller, P. A., and Vousden, K. H. (2013). p53 mutations in cancer. Nat Cell Biol 15, 2-8. Nagata, S. (1997). Apoptosis by death factor. Cell 88, 355-365. Naidu, R., Wahab, N. A., Yadav, M., and Kutty, M. K. (2002). Protein expression and molecular analysis of c-myc gene in primary breast carcinomas using immunohistochemistry and differential polymerase chain reaction. International journal of molecular medicine 9, 189-196. Nemajerova, A., Erster, S., and Moll, U. M. (2005). The post-translational phosphorylation and acetylation modification profile is not the determining factor in targeting endogenous stress-induced p53 to mitochondria. Cell Death Differ 12, 197-200.

Page 97: Apoptosis by c-Myc - Helsinki

Newbold, R. F., and Overell, R. W. (1983). Fibroblast immortality is a prerequisite for transformation by EJ c-Ha-ras oncogene. Nature 304, 648-651. Nicholson, D. W. (1999). Caspase structure, proteolytic substrates, and function during apoptotic cell death. Cell Death Differ 6, 1028-1042. Oakman, C., Viale, G., and Di Leo, A. (2010). Management of triple negative breast cancer. Breast 19, 312-321. Obaya, A. J., Mateyak, M. K., and Sedivy, J. M. (1999). Mysterious liaisons: the relationship between c-Myc and the cell cycle. Oncogene 18, 2934-2941. Okita, K., Ichisaka, T., and Yamanaka, S. (2007). Generation of germline-competent induced pluripotent stem cells. Nature 448, 313-317. Olivier, M., Hollstein, M., and Hainaut, P. (2010). TP53 mutations in human cancers: origins, consequences, and clinical use. Cold Spring Harbor perspectives in biology 2, a001008. Oltersdorf, T., Elmore, S. W., Shoemaker, A. R., Armstrong, R. C., Augeri, D. J., Belli, B. A., Bruncko, M., Deckwerth, T. L., Dinges, J., Hajduk, P. J., et al. (2005). An inhibitor of Bcl-2 family proteins induces regression of solid tumours. Nature 435, 677-681. Ow, Y. P., Green, D. R., Hao, Z., and Mak, T. W. (2008). Cytochrome c: functions beyond respiration. Nature reviews Molecular cell biology 9, 532-542. Pan, G., O'Rourke, K., Chinnaiyan, A. M., Gentz, R., Ebner, R., Ni, J., and Dixit, V. M. (1997). The receptor for the cytotoxic ligand TRAIL. Science 276, 111-113. Park, B. S., Song, Y. S., Yee, S. B., Lee, B. G., Seo, S. Y., Park, Y. C., Kim, J. M., Kim, H. M., and Yoo, Y. H. (2005). Phospho-ser 15-p53 translocates into mitochondria and interacts with Bcl-2 and Bcl-xL in eugenol-induced apoptosis. Apoptosis 10, 193-200. Partanen, J. I., Nieminen, A. I., and Klefstrom, J. (2009). 3D view to tumor suppression: Lkb1, polarity and the arrest of oncogenic c-Myc. Cell Cycle 8, 716-724. Pelengaris, S., Khan, M., and Evan, G. I. (2002). Suppression of Myc-induced apoptosis in beta cells exposes multiple oncogenic properties of Myc and triggers carcinogenic progression. Cell 109, 321-334. Perou, C. M., Sorlie, T., Eisen, M. B., van de Rijn, M., Jeffrey, S. S., Rees, C. A., Pollack, J. R., Ross, D. T., Johnsen, H., Akslen, L. A., et al. (2000). Molecular portraits of human breast tumours. Nature 406, 747-752. Petros, A. M., Gunasekera, A., Xu, N., Olejniczak, E. T., and Fesik, S. W. (2004). Defining the p53 DNA-binding domain/Bcl-x(L)-binding interface using NMR. FEBS Lett 559, 171-174. Pfeifer-Ohlsson, S., Rydnert, J., Goustin, A. S., Larsson, E., Betsholtz, C., and Ohlsson, R. (1985). Cell-type-specific pattern of myc protooncogene expression in developing human embryos. Proc Natl Acad Sci U S A 82, 5050-5054. Pietsch, E. C., Leu, J. I., Frank, A., Dumont, P., George, D. L., and Murphy, M. E. (2007). The tetramerization domain of p53 is required for efficient BAK oligomerization. Cancer Biol Ther 6, 1576-1583. Pietsch, E. C., Perchiniak, E., Canutescu, A. A., Wang, G., Dunbrack, R. L., and Murphy, M. E. (2008). Oligomerization of BAK by p53 utilizes conserved residues of the p53 DNA binding domain. J Biol Chem 283, 21294-21304. Pinder, S. E. (2010). Ductal carcinoma in situ (DCIS): pathological features, differential diagnosis, prognostic factors and specimen evaluation. Modern pathology : an official journal of the United States and Canadian Academy of Pathology, Inc 23 Suppl 2, S8-13. Popescu, N. C., and Zimonjic, D. B. (2002). Chromosome-mediated alterations of the MYC gene in human cancer. J Cell Mol Med 6, 151-159. Pouyssegur, J., Dayan, F., and Mazure, N. M. (2006). Hypoxia signalling in cancer and approaches to enforce tumour regression. Nature 441, 437-443. Prendergast, G. C. (1999). Mechanisms of apoptosis by c-Myc. Oncogene 18, 2967-2987. Pusapati, R. V., Rounbehler, R. J., Hong, S., Powers, J. T., Yan, M., Kiguchi, K., McArthur, M. J., Wong, P. K., and Johnson, D. G. (2006). ATM promotes apoptosis and suppresses tumorigenesis in response to Myc. Proc Natl Acad Sci U S A 103, 1446-1451. Rabbitts, P. H., Watson, J. V., Lamond, A., Forster, A., Stinson, M. A., Evan, G., Fischer, W., Atherton, E., Sheppard, R., and Rabbitts, T. H. (1985). Metabolism of c-myc gene products: c-myc mRNA and protein expression in the cell cycle. Embo J 4, 2009-2015. Reubold, T. F., and Eschenburg, S. (2012). A molecular view on signal transduction by the apoptosome. Cellular signalling 24, 1420-1425. Ricci, J. E., Munoz-Pinedo, C., Fitzgerald, P., Bailly-Maitre, B., Perkins, G. A., Yadava, N., Scheffler, I. E., Ellisman, M. H., and Green, D. R. (2004a). Disruption of mitochondrial function during apoptosis is mediated by caspase cleavage of the p75 subunit of complex I of the electron transport chain. Cell 117, 773-786.

Page 98: Apoptosis by c-Myc - Helsinki

Ricci, M. S., Jin, Z., Dews, M., Yu, D., Thomas-Tikhonenko, A., Dicker, D. T., and El-Deiry, W. S. (2004b). Direct repression of FLIP expression by c-myc is a major determinant of TRAIL sensitivity. Mol Cell Biol 24, 8541-8555. Riedl, S. J., and Shi, Y. (2004). Molecular mechanisms of caspase regulation during apoptosis. Nature reviews Molecular cell biology 5, 897-907. Riley, T., Sontag, E., Chen, P., and Levine, A. (2008). Transcriptional control of human p53-regulated genes. Nature reviews Molecular cell biology 9, 402-412. Robanus-Maandag, E. C., Bosch, C. A., Kristel, P. M., Hart, A. A., Faneyte, I. F., Nederlof, P. M., Peterse, J. L., and van de Vijver, M. J. (2003). Association of C-MYC amplification with progression from the in situ to the invasive stage in C-MYC-amplified breast carcinomas. The Journal of pathology 201, 75-82. Rudner, J., Jendrossek, V., Lauber, K., Daniel, P. T., Wesselborg, S., and Belka, C. (2005). Type I and type II reactions in TRAIL-induced apoptosis -- results from dose-response studies. Oncogene 24, 130-140. Rudolph, B., Saffrich, R., Zwicker, J., Henglein, B., Muller, R., Ansorge, W., and Eilers, M. (1996). Activation of cyclin-dependent kinases by Myc mediates induction of cyclin A, but not apoptosis. Embo J 15, 3065-3076. Saraste, A., and Pulkki, K. (2000). Morphologic and biochemical hallmarks of apoptosis. Cardiovasc Res 45, 528-537. Scaffidi, C., Fulda, S., Srinivasan, A., Friesen, C., Li, F., Tomaselli, K. J., Debatin, K. M., Krammer, P. H., and Peter, M. E. (1998). Two CD95 (APO-1/Fas) signaling pathways. Embo J 17, 1675-1687. Schellenberg, B., Wang, P., Keeble, J. A., Rodriguez-Enriquez, R., Walker, S., Owens, T. W., Foster, F., Tanianis-Hughes, J., Brennan, K., Streuli, C. H., and Gilmore, A. P. (2013). Bax exists in a dynamic equilibrium between the cytosol and mitochondria to control apoptotic priming. Mol Cell 49, 959-971. Schmitz, I., Kirchhoff, S., and Krammer, P. H. (2000). Regulation of death receptor-mediated apoptosis pathways. The international journal of biochemistry & cell biology 32, 1123-1136. Schoenenberger, C. A., Andres, A. C., Groner, B., van der Valk, M., LeMeur, M., and Gerlinger, P. (1988). Targeted c-myc gene expression in mammary glands of transgenic mice induces mammary tumours with constitutive milk protein gene transcription. Embo J 7, 169-175. Schulze, A., and Harris, A. L. (2012). How cancer metabolism is tuned for proliferation and vulnerable to disruption. Nature 491, 364-373. Sears, R., Leone, G., DeGregori, J., and Nevins, J. R. (1999). Ras enhances Myc protein stability. Mol Cell 3, 169-179. Sengupta, S., and Harris, C. C. (2005). p53: traffic cop at the crossroads of DNA repair and recombination. Nature reviews Molecular cell biology 6, 44-55. Seoane, J., Pouponnot, C., Staller, P., Schader, M., Eilers, M., and Massague, J. (2001). TGFbeta influences Myc, Miz-1 and Smad to control the CDK inhibitor p15INK4b. Nat Cell Biol 3, 400-408. Shackelford, D. B., and Shaw, R. J. (2009). The LKB1-AMPK pathway: metabolism and growth control in tumour suppression. Nat Rev Cancer 9, 563-575. Shaw, R. J., Kosmatka, M., Bardeesy, N., Hurley, R. L., Witters, L. A., DePinho, R. A., and Cantley, L. C. (2004). The tumor suppressor LKB1 kinase directly activates AMP-activated kinase and regulates apoptosis in response to energy stress. Proc Natl Acad Sci U S A 101, 3329-3335. Shchors, K., Shchors, E., Rostker, F., Lawlor, E. R., Brown-Swigart, L., and Evan, G. I. (2006). The Myc-dependent angiogenic switch in tumors is mediated by interleukin 1beta. Genes Dev 20, 2527-2538. Sheridan, J. P., Marsters, S. A., Pitti, R. M., Gurney, A., Skubatch, M., Baldwin, D., Ramakrishnan, L., Gray, C. L., Baker, K., Wood, W. I., et al. (1997). Control of TRAIL-induced apoptosis by a family of signaling and decoy receptors. Science 277, 818-821. Sherr, C. J. (2001). The INK4a/ARF network in tumour suppression. Nature reviews Molecular cell biology 2, 731-737. Sherr, C. J. (2004). Principles of tumor suppression. Cell 116, 235-246. Shin, K., Fogg, V. C., and Margolis, B. (2006). Tight junctions and cell polarity. Annual review of cell and developmental biology 22, 207-235. Siegmund, D., Mauri, D., Peters, N., Juo, P., Thome, M., Reichwein, M., Blenis, J., Scheurich, P., Tschopp, J., and Wajant, H. (2001). Fas-associated death domain protein (FADD) and caspase-8 mediate up-regulation of c-Fos by Fas ligand and tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) via a FLICE inhibitory protein (FLIP)-regulated pathway. J Biol Chem 276, 32585-32590. Slamon, D. J., Clark, G. M., Wong, S. G., Levin, W. J., Ullrich, A., and McGuire, W. L. (1987). Human breast cancer: correlation of relapse and survival with amplification of the HER-2/neu oncogene. Science 235, 177-182. Slee, E. A., Keogh, S. A., and Martin, S. J. (2000). Cleavage of BID during cytotoxic drug and UV radiation-induced apoptosis occurs downstream of the point of Bcl-2 action and is catalysed by caspase-3: a

Page 99: Apoptosis by c-Myc - Helsinki

potential feedback loop for amplification of apoptosis-associated mitochondrial cytochrome c release. Cell Death Differ 7, 556-565. Sluss, H. K., Gannon, H., Coles, A. H., Shen, Q., Eischen, C. M., and Jones, S. N. (2010). Phosphorylation of p53 serine 18 upregulates apoptosis to suppress Myc-induced tumorigenesis. Mol Cancer Res 8, 216-222. Sot, B., Freund, S. M., and Fersht, A. R. (2007). Comparative biophysical characterization of p53 with the pro-apoptotic BAK and the anti-apoptotic BCL-xL. J Biol Chem 282, 29193-29200. Soucie, E. L., Annis, M. G., Sedivy, J., Filmus, J., Leber, B., Andrews, D. W., and Penn, L. Z. (2001). Myc potentiates apoptosis by stimulating Bax activity at the mitochondria. Mol Cell Biol 21, 4725-4736. Speidel, D. (2010). Transcription-independent p53 apoptosis: an alternative route to death. Trends in cell biology 20, 14-24. Sporn, M. B. (1996). The war on cancer. Lancet 347, 1377-1381. Srivastava, S., Zou, Z. Q., Pirollo, K., Blattner, W., and Chang, E. H. (1990). Germ-line transmission of a mutated p53 gene in a cancer-prone family with Li-Fraumeni syndrome. Nature 348, 747-749. Staller, P., Peukert, K., Kiermaier, A., Seoane, J., Lukas, J., Karsunky, H., Moroy, T., Bartek, J., Massague, J., Hanel, F., and Eilers, M. (2001). Repression of p15INK4b expression by Myc through association with Miz-1. Nat Cell Biol 3, 392-399. Steinberg, G. R., and Kemp, B. E. (2009). AMPK in Health and Disease. Physiological reviews 89, 1025-1078. Stennicke, H. R., and Salvesen, G. S. (1997). Biochemical characteristics of caspases-3, -6, -7, and -8. J Biol Chem 272, 25719-25723. Stewart, T. A., Pattengale, P. K., and Leder, P. (1984). Spontaneous mammary adenocarcinomas in transgenic mice that carry and express MTV/myc fusion genes. Cell 38, 627-637. Strasser, A., Cory, S., and Adams, J. M. (2011). Deciphering the rules of programmed cell death to improve therapy of cancer and other diseases. EMBO J 30, 3667-3683. Strasser, A., Harris, A. W., Bath, M. L., and Cory, S. (1990). Novel primitive lymphoid tumours induced in transgenic mice by cooperation between myc and bcl-2. Nature 348, 331-333. Suen, D. F., Norris, K. L., and Youle, R. J. (2008). Mitochondrial dynamics and apoptosis. Genes Dev 22, 1577-1590. Susin, S. A., Lorenzo, H. K., Zamzami, N., Marzo, I., Snow, B. E., Brothers, G. M., Mangion, J., Jacotot, E., Costantini, P., Loeffler, M., et al. (1999). Molecular characterization of mitochondrial apoptosis-inducing factor. Nature 397, 441-446. Suzuki, H. I., and Miyazono, K. (2013). p53 actions on microRNA expression and maturation pathway. Methods Mol Biol 962, 165-181. Tait, S. W., and Green, D. R. (2010). Mitochondria and cell death: outer membrane permeabilization and beyond. Nature reviews Molecular cell biology 11, 621-632. Tal, M., King, C. R., Kraus, M. H., Ullrich, A., Schlessinger, J., and Givol, D. (1987). Human HER2 (neu) promoter: evidence for multiple mechanisms for transcriptional initiation. Mol Cell Biol 7, 2597-2601. Taylor, R. C., Cullen, S. P., and Martin, S. J. (2008). Apoptosis: controlled demolition at the cellular level. Nature reviews Molecular cell biology 9, 231-241. Tiainen, M., Vaahtomeri, K., Ylikorkala, A., and Makela, T. P. (2002). Growth arrest by the LKB1 tumor suppressor: induction of p21(WAF1/CIP1). Human molecular genetics 11, 1497-1504. Tiainen, M., Ylikorkala, A., and Makela, T. P. (1999). Growth suppression by Lkb1 is mediated by a G(1) cell cycle arrest. Proc Natl Acad Sci U S A 96, 9248-9251. Tinel, A., and Tschopp, J. (2004). The PIDDosome, a protein complex implicated in activation of caspase-2 in response to genotoxic stress. Science 304, 843-846. Tomiyama, A., Serizawa, S., Tachibana, K., Sakurada, K., Samejima, H., Kuchino, Y., and Kitanaka, C. (2006). Critical role for mitochondrial oxidative phosphorylation in the activation of tumor suppressors Bax and Bak. J Natl Cancer Inst 98, 1462-1473. Tsujimoto, Y., Finger, L. R., Yunis, J., Nowell, P. C., and Croce, C. M. (1984). Cloning of the chromosome breakpoint of neoplastic B cells with the t(14;18) chromosome translocation. Science 226, 1097-1099. Uren, R. T., Dewson, G., Chen, L., Coyne, S. C., Huang, D. C., Adams, J. M., and Kluck, R. M. (2007). Mitochondrial permeabilization relies on BH3 ligands engaging multiple prosurvival Bcl-2 relatives, not Bak. J Cell Biol 177, 277-287. Vaahtomeri, K., and Makela, T. P. (2011). Molecular mechanisms of tumor suppression by LKB1. FEBS Lett 585, 944-951. van Delft, M. F., Wei, A. H., Mason, K. D., Vandenberg, C. J., Chen, L., Czabotar, P. E., Willis, S. N., Scott, C. L., Day, C. L., Cory, S., et al. (2006). The BH3 mimetic ABT-737 targets selective Bcl-2 proteins and efficiently induces apoptosis via Bak/Bax if Mcl-1 is neutralized. Cancer Cell 10, 389-399.

Page 100: Apoptosis by c-Myc - Helsinki

Vander Heiden, M. G., Cantley, L. C., and Thompson, C. B. (2009). Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science 324, 1029-1033. Vaseva, A. V., Marchenko, N. D., and Moll, U. M. (2009). The transcription-independent mitochondrial p53 program is a major contributor to nutlin-induced apoptosis in tumor cells. Cell Cycle 8, 1711-1719. Vassilev, L. T., Vu, B. T., Graves, B., Carvajal, D., Podlaski, F., Filipovic, Z., Kong, N., Kammlott, U., Lukacs, C., Klein, C., et al. (2004). In vivo activation of the p53 pathway by small-molecule antagonists of MDM2. Science 303, 844-848. Vaux, D. L., and Silke, J. (2005). IAPs, RINGs and ubiquitylation. Nature reviews Molecular cell biology 6, 287-297. Vazquez, A., Markert, E. K., and Oltvai, Z. N. (2011). Serine biosynthesis with one carbon catabolism and the glycine cleavage system represents a novel pathway for ATP generation. PLoS One 6, e25881. Veikkola, T., and Alitalo, K. (1999). VEGFs, receptors and angiogenesis. Semin Cancer Biol 9, 211-220. Vennstrom, B., Sheiness, D., Zabielski, J., and Bishop, J. M. (1982). Isolation and characterization of c-myc, a cellular homolog of the oncogene (v-myc) of avian myelocytomatosis virus strain 29. Journal of virology 42, 773-779. Verhagen, A. M., Ekert, P. G., Pakusch, M., Silke, J., Connolly, L. M., Reid, G. E., Moritz, R. L., Simpson, R. J., and Vaux, D. L. (2000). Identification of DIABLO, a mammalian protein that promotes apoptosis by binding to and antagonizing IAP proteins. Cell 102, 43-53. Verhagen, A. M., Silke, J., Ekert, P. G., Pakusch, M., Kaufmann, H., Connolly, L. M., Day, C. L., Tikoo, A., Burke, R., Wrobel, C., et al. (2002). HtrA2 promotes cell death through its serine protease activity and its ability to antagonize inhibitor of apoptosis proteins. J Biol Chem 277, 445-454. Vogelstein, B., Lane, D., and Levine, A. J. (2000). Surfing the p53 network. Nature 408, 307-310. Wade, M., Wong, E. T., Tang, M., Stommel, J. M., and Wahl, G. M. (2006). Hdmx modulates the outcome of p53 activation in human tumor cells. J Biol Chem 281, 33036-33044. Walczak, H., Miller, R. E., Ariail, K., Gliniak, B., Griffith, T. S., Kubin, M., Chin, W., Jones, J., Woodward, A., Le, T., et al. (1999). Tumoricidal activity of tumor necrosis factor-related apoptosis-inducing ligand in vivo. Nat Med 5, 157-163. Wang, K., Yin, X. M., Chao, D. T., Milliman, C. L., and Korsmeyer, S. J. (1996). BID: a novel BH3 domain-only death agonist. Genes Dev 10, 2859-2869. Warburg, O., Wind, F., and Negelein, E. (1927). The Metabolism of Tumors in the Body. The Journal of general physiology 8, 519-530. Ward, P. S., and Thompson, C. B. (2012). Metabolic reprogramming: a cancer hallmark even warburg did not anticipate. Cancer Cell 21, 297-308. Watson, C. J., and Khaled, W. T. (2008). Mammary development in the embryo and adult: a journey of morphogenesis and commitment. Development 135, 995-1003. Wei, M. C., Lindsten, T., Mootha, V. K., Weiler, S., Gross, A., Ashiya, M., Thompson, C. B., and Korsmeyer, S. J. (2000). tBID, a membrane-targeted death ligand, oligomerizes BAK to release cytochrome c. Genes Dev 14, 2060-2071. Wei, M. C., Zong, W. X., Cheng, E. H., Lindsten, T., Panoutsakopoulou, V., Ross, A. J., Roth, K. A., MacGregor, G. R., Thompson, C. B., and Korsmeyer, S. J. (2001). Proapoptotic BAX and BAK: a requisite gateway to mitochondrial dysfunction and death. Science 292, 727-730. Weinberg, R. A. (1995). The retinoblastoma protein and cell cycle control. Cell 81, 323-330. Weinberg, R. A. (2007). The Biology of Cancer. Weng, C., Li, Y., Xu, D., Shi, Y., and Tang, H. (2005). Specific cleavage of Mcl-1 by caspase-3 in tumor necrosis factor-related apoptosis-inducing ligand (TRAIL)-induced apoptosis in Jurkat leukemia T cells. J Biol Chem 280, 10491-10500. Werb, Z. (1997). ECM and cell surface proteolysis: regulating cellular ecology. Cell 91, 439-442. Willis, S. N., Chen, L., Dewson, G., Wei, A., Naik, E., Fletcher, J. I., Adams, J. M., and Huang, D. C. (2005). Proapoptotic Bak is sequestered by Mcl-1 and Bcl-xL, but not Bcl-2, until displaced by BH3-only proteins. Genes Dev 19, 1294-1305. Willis, S. N., Fletcher, J. I., Kaufmann, T., van Delft, M. F., Chen, L., Czabotar, P. E., Ierino, H., Lee, E. F., Fairlie, W. D., Bouillet, P., et al. (2007). Apoptosis initiated when BH3 ligands engage multiple Bcl-2 homologs, not Bax or Bak. Science 315, 856-859. Wilson, A., Murphy, M. J., Oskarsson, T., Kaloulis, K., Bettess, M. D., Oser, G. M., Pasche, A. C., Knabenhans, C., Macdonald, H. R., and Trumpp, A. (2004). c-Myc controls the balance between hematopoietic stem cell self-renewal and differentiation. Genes Dev 18, 2747-2763. Wilson, W. H., O'Connor, O. A., Czuczman, M. S., LaCasce, A. S., Gerecitano, J. F., Leonard, J. P., Tulpule, A., Dunleavy, K., Xiong, H., Chiu, Y. L., et al. (2010). Navitoclax, a targeted high-affinity inhibitor of

Page 101: Apoptosis by c-Myc - Helsinki

BCL-2, in lymphoid malignancies: a phase 1 dose-escalation study of safety, pharmacokinetics, pharmacodynamics, and antitumour activity. The lancet oncology 11, 1149-1159. Wise, D. R., DeBerardinis, R. J., Mancuso, A., Sayed, N., Zhang, X. Y., Pfeiffer, H. K., Nissim, I., Daikhin, E., Yudkoff, M., McMahon, S. B., and Thompson, C. B. (2008). Myc regulates a transcriptional program that stimulates mitochondrial glutaminolysis and leads to glutamine addiction. Proc Natl Acad Sci U S A 105, 18782-18787. Wise, D. R., and Thompson, C. B. (2010). Glutamine addiction: a new therapeutic target in cancer. Trends in biochemical sciences 35, 427-433. Wright, W. E., Pereira-Smith, O. M., and Shay, J. W. (1989). Reversible cellular senescence: implications for immortalization of normal human diploid fibroblasts. Mol Cell Biol 9, 3088-3092. Wyllie, A. H., Kerr, J. F., and Currie, A. R. (1980). Cell death: the significance of apoptosis. Int Rev Cytol 68, 251-306. Xu, H., Tai, J., Ye, H., Kang, C. B., and Yoon, H. S. (2006). The N-terminal domain of tumor suppressor p53 is involved in the molecular interaction with the anti-apoptotic protein Bcl-Xl. Biochem Biophys Res Commun 341, 938-944. Xu, J., Chen, Y., and Olopade, O. I. (2010). MYC and Breast Cancer. Genes & cancer 1, 629-640. Xu, Y. (2003). Regulation of p53 responses by post-translational modifications. Cell Death Differ 10, 400-403. Ylikorkala, A., Avizienyte, E., Tomlinson, I. P., Tiainen, M., Roth, S., Loukola, A., Hemminki, A., Johansson, M., Sistonen, P., Markie, D., et al. (1999). Mutations and impaired function of LKB1 in familial and non-familial Peutz-Jeghers syndrome and a sporadic testicular cancer. Human molecular genetics 8, 45-51. Youle, R. J., and Strasser, A. (2008). The BCL-2 protein family: opposing activities that mediate cell death. Nature reviews Molecular cell biology 9, 47-59. Yuneva, M., Zamboni, N., Oefner, P., Sachidanandam, R., and Lazebnik, Y. (2007). Deficiency in glutamine but not glucose induces MYC-dependent apoptosis in human cells. J Cell Biol 178, 93-105. Zha, J., Harada, H., Yang, E., Jockel, J., and Korsmeyer, S. J. (1996). Serine phosphorylation of death agonist BAD in response to survival factor results in binding to 14-3-3 not BCL-X(L). Cell 87, 619-628. Zhan, L., Rosenberg, A., Bergami, K. C., Yu, M., Xuan, Z., Jaffe, A. B., Allred, C., and Muthuswamy, S. K. (2008). Deregulation of scribble promotes mammary tumorigenesis and reveals a role for cell polarity in carcinoma. Cell 135, 865-878. Zilfou, J. T., and Lowe, S. W. (2009). Tumor suppressive functions of p53. Cold Spring Harbor perspectives in biology 1, a001883. Zindy, F., Eischen, C. M., Randle, D. H., Kamijo, T., Cleveland, J. L., Sherr, C. J., and Roussel, M. F. (1998). Myc signaling via the ARF tumor suppressor regulates p53-dependent apoptosis and immortalization. Genes Dev 12, 2424-2433. Zorbas, H., Luxford, K., Evans, A., and Villanueva, E. V. (2004). Re: Trends in the treatment of ductal carcinoma in situ of the breast. J Natl Cancer Inst 96, 1258; author reply 1259-1260. Zou, H., Li, Y., Liu, X., and Wang, X. (1999). An APAF-1.cytochrome c multimeric complex is a functional apoptosome that activates procaspase-9. J Biol Chem 274, 11549-11556.

Page 102: Apoptosis by c-Myc - Helsinki