ArticlePDF AvailableLiterature Review

Multiple cell death pathways as regulators of tumour initiation and progression

Authors:
  • Danish Cancer Institute

Abstract and Figures

Acquired defects in signalling pathways leading to programmed cell death (PCD) are among the major hallmarks of cancer. Although focus has been on caspase-dependent apoptotic death pathways, evidence is now accumulating that nonapoptotic PCD also can form an important barrier against tumour initiation and progression. Akin to the earlier landmark discoveries that lead to the identification of the major cancer-related proteins like p53, c-Myc and Bcl-2 as controllers of spontaneous and therapy-induced apoptosis, numerous proteins with properties of tumour suppressors and oncoproteins have recently been identified as key regulators of alternative death programmes. The emerging data on the molecular mechanisms regulating nonapoptotic PCD may have potent therapeutic consequences.
Content may be subject to copyright.
Multiple cell death pathways as regulators of tumour initiation and
progression
Marja Ja
¨
a
¨
ttela
¨
*
,1
1
Apoptosis Laboratory, Institute of Cancer Biology, Danish Cancer Society, Strandboulevarden 49, DK-2100 Copenhagen, Denmark
Acquired defects in signalling pathways leading to
programmed cell death (PCD) are among the major
hallmarks of cancer. Although focus has been on caspase-
dependent apoptotic death pathways, evidence is now
accumulating that nonapoptotic PCD also can form an
important barrier against tumour initiation and progres-
sion. Akin to the earlier landmark discoveries that lead to
the identification of the major cancer-related proteins like
p53, c-Myc and Bcl-2 as controllers of spontaneous and
therapy-induced apoptosis, numerous proteins with prop-
erties of tumour suppressors and oncoproteins have
recently been identified as key regulators of alternative
death programmes. The emerging data on the molecular
mechanisms regulating nonapoptotic PCD may have
potent therapeutic consequences.
Oncogene (2004) 23, 2746–2756. doi:10.1038/sj.onc.1207513
Keywords: cancer; caspase independent; cell death;
lysosome; transformation
Introduction
Cancer is a disease characterized by an imbalance
between cell division and cell death (Hanahan and
Weinberg, 2000). Apoptosis is the best-defined cell death
programme counteracting tumour growth. It is char-
acterized by the activation of a specific family of cysteine
proteases, the caspases, followed by a series of caspase-
mediated morphological changes such as the shrinkage
of the cell, the condensation of the chromatin and the
disintegration of the cell into small fragments that can
be engulfed by nearby cells without inciting inflamma-
tion (Kerr et al., 1972; Strasser et al., 2000; Ferri and
Kroemer, 2001; Kaufmann and Hengartner, 2001). With
respect to the ability of some cells to survive caspase
activation, it would be dangerous for the organism to
depend on a single protease family for the clearance of
unwanted and potentially harmful cells (Leist and
Ja
¨
a
¨
ttela
¨
, 2001). Indeed, accumulating data now show
that programmed cell death (PCD) can occur in the
complete absence of caspases, and evidence for death
pathways where noncaspase proteases and other death
effectors function as executioners is emerging (Ferri and
Kroemer, 2001; Leist and Ja
¨
a
¨
ttela
¨
, 2001; Lockshin and
Zakeri, 2002). Experiments employing cancer cells with
defective apoptosis machinery have revealed that most
caspase-activating apoptotic stimuli, including onco-
genes, p53, DNA-damaging drugs, proapoptotic Bcl-2
family members, cytotoxic lymphocytes and in some
cases even death receptors, do not require known
caspases for PCD to occur (Leist and Ja
¨
a
¨
ttela
¨
, 2001;
Mathiasen and Ja
¨
a
¨
ttela
¨
, 2002). Thus, in most death
pathways the role of caspases is rather at the level of
shaping the destruction of the cell that is already
committed to die than at the level of the decision to die.
In the following, I will first describe various types of PCD
and their control at molecular and organelle level and
then discuss the numerous links that connect transforma-
tion and tumour progression to the control of PCD.
Classification of PCD according to the nuclear
morphology
The foremost criterion for PCD (or active cell death)
that distinguishes it from accidental necrosis is the
participation of active cellular processes that can be
intercepted by interfering with intracellular signalling
(Leist and Ja
¨
a
¨
ttela
¨
, 2001). The unclear definition of
caspase-independent death pathways has been the major
obstacle to their better understanding. Since the precise
characterization of biochemical check points controlling
caspase-independent PCD is still awaiting, the nuclear
morphology of dying cells (Leist and Ja
¨
a
¨
ttela
¨
, 2001) has
been used as an alternative basis for the classification of
PCD (Figure 1). According to the nuclear morphology,
PCD can be divided into three subclasses: (i) classic
apoptosis characterized by chromatin condensed to
compact and almost geometric figures (stage 2 chroma-
tin condensation), (ii) apoptosis-like PCD with less
compact, lumpy chromatin masses (stage 1 chromatin
condensation) and (iii) necrosis-like PCD that occurs
either in the complete absence of chromatin condensa-
tion or at best with chromatin clustering to form loose
speckles (Leist and Ja
¨
a
¨
ttela
¨
, 2001). So far, nearly all
examples of classic apoptosis appear caspase dependent,
caspases playing an active role in the activation of DNA
fragmentation factor 45 (DFF45) that is responsible for
the compact chromatin condensation (Sakahira et al.,
1998). Theoretically, stage 2 chromatin condensation
*Correspondence: M Ja
¨
a
¨
ttela
¨
; E-mail: mhj@biobase.dk
Oncogene (2004) 23, 2746–2756
&
2004 Nature Publishing Group
All rights reserved 0950-9232/04 $25.00
www.nature.com/onc
could, however, occur caspase independently by
endonuclease G (EndoG), a mitochondrial DNase that
is released in response to various apoptotic stimuli (Li
et al., 2001). Most published forms of ‘caspase-
independent apoptosis’ fall into the category of apop-
tosis-like PCD even though some of the classic ‘caspase-
dependent apoptosis’ models also display this morphol-
ogy, for example due to the lack of DFF45 expression.
Apoptosis-inducing factor (AIF) released from the
mitochondria during various forms of PCD (Susin
et al., 1999), calcium-regulated cytosolic cysteine pro-
teases calpains, cathepsin B, a lysosomal cysteine
protease that translocates into the cytosol and nucleus
in response to many cellular stresses (Vancompernolle
et al., 1998), and
L-DNase II derived from the leucocyte
elastase inhibitor by post-translational modification
induced by low pH or digestion with elastase (Torriglia
et al., 1998), are among the effector molecules that can
trigger the stage 1 chromatin condensation. Autophagic
degeneration (also called type II PCD) (Chi et al., 1999)
and death receptor-induced necrosis (Vercammen et al.,
1998a) lack chromatin condensation and can thus be
classified as necrosis-like PCD. Owing to the enormous
overlap and shared signalling pathways between the
different death programmes, cells dying by apoptosis-
and necrosis-like PCD can display any degree and
combination of other apoptotic features including
phosphatidylserine exposure, cytoplasmic shrinkage,
zeiosis, formation of apoptotic bodies, and even the
activation of caspases (Leist and Ja
¨
a
¨
ttela
¨
, 2001). It
should also be noted that a single stimulus often triggers
several distinct death programmes concurrently. Nor-
mally, only the fastest and most effective death pathway
is evident, but one cell may also display characteristics
of several death programmes simultaneously (Bursch
et al., 1996).
Classification of PCD according to the involvement of
cellular compartments
The classification of PCD based on the nuclear
morphology does not take into the account the death
signalling pathways involved. Therefore, attempts to
sort PCD according to the cellular compartments
involved in the process (mitochondria, lysosomes,
plasma membrane receptors, nuclei, cytoskeleton and
endoplasmic reticulum) may provide a more informative
basis for the taxonomy (Ferri and Kroemer, 2001). It
does, however, not offer a simple solution, because most
death pathways depend on input from several parts of
the cell, the major one being the mitochondrion that is
essential for the vast majority of death pathways
(Figure 1). As mitochondrial membrane permeabiliza-
tion (MMP) is the event that defines the point of no
return in most PCD models, its control and conse-
quences are described first followed by the depiction of
the involvement of other cellular compartments to
MMP-dependent as well as MMP-independent PCD.
Control of MMP
The pathways upstream of MMP are numerous and
with only few exceptions tightly controlled by the
members of the Bcl-2 family (Coultas and Strasser,
2003) (Figure 1). Bax and Bak, the ‘multidomain’
proapoptotic Bcl-2 family members, are crucial pore-
forming molecules that trigger MMP and the release
of death-inducing molecules from the mitochondrial
intermembrane space. Bax/Bak induce initially outer
Figure 1 The control and consequences of MMP. Pore-forming
proteins Bax and/or Bak can trigger MMP following activation by
BH3-only proteins or cleavage. Caspase-8 and cysteine cathepsin
can cleave and activate a BH3-only protein, Bid, disruption of the
cytoskeleton leads to the release of BH3-only proteins Bim and
Bmf, activation of the JNK pathway increases the expression Bmf
and Hrk, DNA damage induces a p53-mediated transcription of
genes encoding Bax, BH3-domain only proteins (Noxa and
PUMA) as well as proteins involved in ROS generation, ER stress
results in the release of Ca
2 þ
, which may cause direct mitochondrial
damage or activate Bax via calpain-mediated cleavage and
hypoxia-activated transcription factor HIF-1a induces the expres-
sion of BNIP3, whose stability is increased by acidosis, and
activation of CD47 releases BNIP3 from the receptor allowing it to
translocate to the mitochondrial membrane. Furthermore, various
death stimuli trigger the production of lipid second messengers that
are involved in MMP and mitochondrial damage. Mitochondrial
damage leads to the release of numerous mitochondrial proteins
that trigger the execution of PCD. Cytochrome c, Smac/Diablo and
Omi/Htra2 triggers caspase activation and classic apoptosis. AIF
triggers a caspase-independent death pathway culminating in DNA
fragmentation and stage 1 chromatin condensation characteristic
of apoptosis-like PCD. Endonuclease G cleaves DNA and induces
stage 2 chromatin condensation. The serine protease activity of
Omi/Htra2 mediates caspase-independent cellular rounding and
shrinkage without changes in the nuclear morphology. Ca
2 þ
and
ROS can lead into severe mitochondrial dysfunction and necrosis-
like PCD with or without autophagy. The schematic drawings in
the bottom of the figure provide a general guideline to differentiate
between the different PCD forms according to the nuclear
morphology. Caspase-dependent chromatin compaction and frag-
mentation to crescent- or spherical-shaped masses at the nuclear
periphery is shown in the middle. Caspase-independent chromatin
margination triggered directly by microinjection of AIF or in a
number of other models of apoptosis-like death is shown at right.
No chromatin condensation, but sometimes passive clumping and
dissolution of nucleoli is characteristic of necrotic PCD
Cell death pathways as regulators of tumour initiation and progression
MJa¨a¨ttela¨
2747
Oncogene
membrane permeability and leave the inner membrane,
the protein import function and the ultrastructure of
mitochondria intact (Von Ahsen et al., 2000). Bax/Bak
can be activated transcriptionally or by conformational
change induced by cleavage or binding to BH3-only
proteins (a subgroup of proapoptotic Bcl-2 proteins that
share homology to Bcl-2 only in the Bcl-2 homology
domain 3). Antiapoptotic Bcl-2 proteins (e.g. Bcl-2 and
Bcl-X
L
) oppose MMP presumably by heterodimeriza-
tion with Bax-like proteins. In contrast, proapoptotic
‘BH3-only’ proteins (Bad, Bid, Bim, Bmf, Noxa,
PUMA, BNIP3, etc.) either oppose the inhibitory effect
of Bcl-2-like proteins via BH3-mediated binding or
activate Bax-like proteins by direct binding. BNIP3, a
death-promoting ‘BH3-only’ protein differs from the
other proteins of this class. It can integrate into the
mitochondrial outer membrane in a BH3-domain- and
presumably also Bax/Bak-independent manner to trig-
ger the loss of mitochondrial transmembrane potential,
reactive oxygen species (ROS) generation and PCD with
characteristics of autophagic degeneration (Vande Velde
et al., 2000). Similarly, overexpression of Hspin1,
another Bcl-2 binding transmembrane protein with no
BH3 domain, causes upregulation of a lysosomal
protease cathepsin D and autophagy without release of
proapoptotic proteins from the mitochondrial inter-
membrane space (Yanagisawa et al., 2003). Further-
more, increase in cytosolic calcium and arachidonic acid
may induce mitochondrial damage in a Bax/Bak-
independent manner (Scorrano et al., 2001; Scorrano
et al., 2003).
Consequences of MMP
MMP limited to the mitochondrial outer membrane is
sufficient for the release of death-promoting molecules
from the mitochondrial intermembrane space, whereas
permeabilization of both membranes normally results in
necrosis-like PCD characterized by the loss of mito-
chondrial transmembrane potential and major meta-
bolic disturbances due to the reduced ATP generation
and increased production of ROS (Figure 1). The
classical apoptosis pathway is initiated by the release
of cytochrome c (Li et al., 1997). It prompts the ATP-
dependent assembly of the apoptosome complex con-
sisting of cytochrome c, apoptotic protease-activating
factor 1 and caspase-9, which forms the template for
efficient caspase processing and activation. As a further
safe-guard mechanism, caspase-inhibitory factors (pro-
teins of inhibitor of apoptosis protein family) have to be
removed by additional proteins (Smac/Diablo or Omi/
htra2) released from mitochondria before the execution
caspases can become fully active and bring forth the
typically apoptotic morphology best characterized by
the compact condensation of the chromatin (Du et al.,
2000; Suzuki et al., 2001).
Another distinct pathway following MMP is mediated
by AIF (Susin et al., 1999). Once released from the
mitochondria, AIF translocates into the nucleus where it
induces caspase-independent type 1 chromatin conden-
sation by an as yet undefined mechanism. EndoG and a
serine protease Omi/Htra2 may also contribute to the
caspase-independent death signalling downstream of
MMP. Extramitochondrial expression of Omi triggers
necrosis-like PCD that is dependent on Omi’s serine
protease activity (Suzuki et al., 2001), and EndoG can
cause caspase-independent DNA fragmentation in iso-
lated nuclei (Li et al., 2001). Caenorhabditis elegans AIF
ortholog (wah-1) associates and cooperates with the C.
elegans EndoG ortholog (CPS-6) to promote DNA
degradation and apoptosis (Wang et al., 2002). Whether
EndoG and AIF also define a single MMP-initiated
DNA degradation pathway in mammalian cells remains
to be studied. Whereas there is emerging evidence
suggesting that AIF may serve as a safe-guard death
executioner in some cancer cells with faulty caspase
activation in vitro as well as in c-Myc-induced mammary
tumours in vivo (Joseph et al., 2002; Alonso et al., 2003;
Liao and Dickson, 2003), the roles of EndoG and Omi/
Htra2 in cancer cell death await to be elucidated.
Instead of the relatively rapid death pathways
described above, mitochondrial damage may also lead
to a slow sequestration of the damaged mitochondria in
autophagic vacuoles (Tolkovsky et al., 2002). If the
damage is limited, autophagy may serve as a rescue
mechanism inhibiting the further release of the death-
promoting mitochondrial proteins to the cytosol. How-
ever, in the case of more severe or continuous damage,
autophagy leads to metabolic death after a lag phase of
several days.
In the majority of cell death models, the master
controllers of PCD operate at the mitochondrial level,
whereas the decision concerning the nature of the death
is taken after this step. There are, however, certain cases
where mitochondria may not hold a regulatory role, but
instead organelles like lysosomes and endoplasmic
reticulum (ER) might be in charge of the execution of
the cell (Chi et al., 1999; Elliott et al., 2000; Nylandsted
et al., 2000; Sperandio et al., 2000; Inbal et al., 2002;
Nylandsted et al., 2002).
Receptor-mediated PCD comes in many shapes
Plasma membrane plays an important role in the life-
death decisions of the cell via its numerous receptors
that promote either survival or PCD. Death receptors of
tumour necrosis factor receptor (TNFR) superfamily,
for example TNFR1 and Fas (also known as CD95) and
receptors for TNF-related apoptosis-inducing ligand
(TRAIL), trigger almost the whole spectrum of different
death pathways either dependent or independent of
MMP and caspases and provide thus an exceptional
example of the complexity of PCD signalling (Ashke-
nazi, 2002; Ja
¨
a
¨
ttela
¨
and Tschopp, 2003). The basic
signalling unit of DRs consists of three receptors, which
are brought together and actived upon binding to the
trimeric ligand. The receptors then assemble a death-
inducing signalling complex. Through a death domain-
mediated binding of the adaptor proteins TRADD
(TNF-R1) and FADD (all death receptors), death
receptors recruit and activate the apoptosis-initiating
proteases caspase-8 and/or caspase-10 that induce
Cell death pathways as regulators of tumour initiation and progression
MJa¨a¨ttela¨
2748
Oncogene
apoptosis either by direct activation of effector caspases
or via a Bax/Bak-dependent MMP triggered by caspase-
8-mediated cleavage of Bid (Luo et al., 1998; Scaffidi
et al., 1998). Another type of signalling complex induced
by death receptors involves the kinase receptor-asso-
ciated protein 1 (RIP1) and the caspase-8 inhibitor
FLIP, which, via their interaction with TNFR-asso-
ciated factors (TRAFs) links death receptors to the
signalling cascades activating the NF-kB transcription
factor and mitogen-activated protein kinases (Thome
and Tschopp, 2001).
Embryonic fibroblasts or thymocytes from mice
deficient in FADD and caspase-8 are resistant to FasL
and TNF (Varfolomeev et al., 1998; Yeh et al., 1998;
Zhang et al., 1998), revealing the central role of caspase
activation and thus apoptosis in death receptor signal-
ling. A number of recent reports show however that
FasL, TNF and TRAIL cause cell death even in the
absence of caspase activation in other cell types. The
nonapoptotic caspase-independent PCD may even be
predominant, for instance in the L929 fibrosarcoma cell
line (Vercammen et al., 1998a, b). These cells respond to
TNF with necrosis-like PCD without the involvement of
caspases, despite the fact that Fas-stimulation leads to
caspase-dependent apoptosis. When caspases are inhib-
ited by various means, the cells still die upon Fas
stimulation, albeit with reduced efficacy, delayed ki-
netics and necrosis-like morphology. Thus, within the
same cell line both death pathways are available to
death receptors but are differentially engaged depending
on the nature of the incoming signal. Since the presence
of caspase inhibitors in L929 cells results not only in the
appearance of a Fas-triggered necrosis, but also in a
1000-fold sensitization towards TNF-induced necrosis,
caspases may be implicated in an antinecrotic pathway.
The DR-induced caspase-independent pathway is not
restricted to the L929 cells. In the U937 and Jurkat
leukaemia, HT29 colon carcinoma, NIH3T3 fibroblast
and WEHI-S fibrosarcoma cells, caspase-independent
PCD with either necrosis- or apoptosis-like morphology
can be engaged by TNF, FasL and/or TRAIL (Khwaja
and Tatton, 1999; Holler et al., 2000; Luschen et al.,
2000; Foghsgaard et al., 2001; Wilson and Browning,
2002). Our understanding of the signalling pathway
leading to DR-induced nonapoptotic PCD and the
mechanisms governing the decision points between the
caspase-dependent and -independent cell death is as yet
limited. FasL-induced necrotic PCD is absent in T cells
that are deficient in FADD, indicating that the presence
of FADD is crucial for both necrotic and apoptotic
PCD induced by FasL (Holler et al., 2000). Another
protein involved in the necrotic signalling cascade is
RIP1. Jurkat cells deficient in RIP1 are completely
resistant to cell death even at high concentrations of
FasL, provided that caspases are inhibited (Holler et al.,
2000). Unlike RIP1’s function in the recruitment of the
activation of NF-kB (Kelliher et al., 1998), the kinase
activity of RIP1 is indispensable for its death-promoting
activity (Holler et al., 2000). This suggests that RIP1
phosphorylates and thereby regulates a key player of the
nonapoptotic cell death programme. Fas can also trigger
caspase-independent apoptosis-like PCD. In immorta-
lized epithelial cells, activated Fas recruits Daxx from
the nucleus to the receptor complex and triggers its
binding with apoptosis signal-regulating kinase 1
(Ask1). In addition to a caspase-dependent proapopto-
tic function that depends on its kinase activity, Ask1
possesses a caspase-independent killing function that is
independent of its kinase activity and is activated by
interaction with Daxx (Charette et al., 2001). The
relationship between RIP1-mediated and Daxx/Ask1-
mediated death pathways remains to be studied.
The caspase-independent PCD triggered by TNF is
mechanistically similar but not identical to that trig-
gered by Fas. While the kinase activity of RIP1 is also
obligatory, FADD appears to be dispensable (Holler
et al., 2000). Furthermore, pharmaceutical inhibition
and antisense cDNA-mediated depletion of the lysoso-
mal cysteine protease cathepsin B confers almost
complete protection against TNF (and TRAIL)-induced
apoptosis-like caspase-independent PCD in WEHI-S
cells (Foghsgaard et al., 2001). Consistent with the
prominent role of cathepsin B in this pathway,
immortalized murine embryonic fibroblasts (iMEFs)
deficient for cathepsin B are highly resistant to TNF
(Foghsgaard et al., 2002). Interestingly, cathepsin B-
deficient primary MEFs display similar sensitivity to
TNF as the wild-type MEFs (Foghsgaard et al., 2001),
but upon cellular immortalization MEFs become over
1000-fold sensitized to TNF-induced apoptosis-like
PCD that is cathepsin B-dependent and caspase-
independent. This sensitization is further increased upon
Ras- and Src-mediated transformation of iMEFs (N
Fehrenbacher and MJ, unpublished). Whether cathepsin
B is on the same pathway as RIP1 is still an open
question. This is, however, supported by data showing
that the treatment of Jurkat cells with TLCK, a potent
inhibitor of serine and cysteine proteases including
cathepsin B, inhibits RIP1-dependent necrosis-like PCD
induced by TNF and FasL (Holler et al., 2000).
There is ample evidence that excess ROS formation is
involved in the nonapoptotic PCD induced by death
receptors. Inhibition of caspases (which sensitizes to
necrosis) results in increased ROS formation and the
addition of butylated hydroxyanisole (BHA), an oxygen
radical scavenger, hinders both TNF- and FasL-induced
necrosis-like PCD in L929 cells and TNF-induced
apoptosis-like PCD in WEHI-S cells and iMEFs
(Vercammen et al., 1998b; Foghsgaard et al., 2002).
TNF-induced ROS has been proposed to be a result of
enhanced electron flow through the mitochondrial
electron transport chain (Schulze-Osthoff et al., 1992).
Other possible intracellular sources of ROS include
lysosomes and phospholipase A2 (PLA2). Interestingly,
the TNF-induced activation of PLA2 is dependent on
cathepsin B (Foghsgaard et al., 2002).
In addition to the death pathways described above,
TNF-R1 triggers caspase-independent autophagy and
membrane blebbing mediated by death-associated pro-
tein kinase (DAPK) and DAPK-related protein kinase 1
(DRP-1) in cells displaying caspase-mediated and
DAPK/DRP1-independent nuclear fragmentation
Cell death pathways as regulators of tumour initiation and progression
MJa¨a¨ttela¨
2749
Oncogene
(Inbal et al., 2002). Interestingly, inhibition of caspases
but not DAPK/DRP1 could rescue cells from death. In
fact DAPK/DRP1 inhibition increased the amount of
cells with condensed and fragmented nuclei, suggesting
that autophagy may function as a rescue mechanism
against caspase-mediated apoptosis in this model system.
Other cell surface receptors that trigger alternative
death pathways in a cancer-specific manner include
CD20, CD47, CD99 and insulin-like growth factor 1
receptor (IGF1R). Rituximab is an anti-CD20 antibody
successfully used for the treatment of lymphoma. In
vitro, it induces a mitochondria- and caspase-indepen-
dent exposure of phosphatidylserine and caspase-inde-
pendent loss of mitochondrial membrane potential,
suggesting that its efficacy may be due to the ability to
kill lymphoma cells in spite of defective caspase
activation or overexpression of Bcl-2-like proteins
(Chan et al., 2003). Activation of CD47, a receptor for
thrombospondin, initiates necrosis-like PCD in B-cell
chronic lymphoma cells (Mateo et al., 1999). Also,
resting B and T cells are affected but to a lesser degree
suggesting that therapies targeting CD47 may prove
useful in the clinic (Mateo et al., 2002; Lamy et al.,
2003). CD47-initiated death is characterized by early
mitochondrial dysfunction, suppression of mitochon-
drial membrane potential and generation of ROS
without signs of cytochrome c release or caspase
activation. BNIP3, the Bcl-2 family protein mediating
autophagy (Vande Velde et al., 2000), may be the
effector molecule in this pathway (Lamy et al., 2003). It
binds to the unoccupied receptor, but translocates to the
mitochondrial membrane upon ligand or antibody
binding. Furthermore, downregulation of BNIP3 with
antisense oligonucletides partially attenuates CD47-
induced necrotic death. Antibodies to yet another
lymphoid receptor, CD99 triggers a rapid and profound
apoptosis-like PCD in transformed but not normal T
cells by an as yet unidentified pathway (Pettersen et al.,
2001).
Ligand-activated IGF1R mediates one of the most
potent survival signals for cancer cells (Evan and
Vousden, 2001). Overexpression of an unoccupied
IGF1R or its cytosolic tail triggers, however, necrosis-
like death that is independent of MMP, Apaf-1 and
caspase activity (Sperandio et al., 2000). Surprisingly,
this pathway is dependent on procaspase-9 protein
suggesting that procaspase-9 may posses another enzy-
matic activity or function as an adaptor protein linking
unidentified death effectors to IGF1R in order to trigger
necrosis-like PCD in cells deprived of IGF. Thus,
IGF1R antagonists may turn out to be more successful
in cancer therapy than anticipated. Furthermore, it will
be interesting to study whether other so-called depen-
dence receptors activate a similar pathway. In the case
of ‘deleted in colorectal cancer’ (DCC), an Apaf-1- and
MMP-independent death pathway has already been
described, but contrary to IGF1R, DCC-induced
apoptosis is dependent on caspase activity (Forcet
et al., 2001). Activation of this pathway may, however,
prove useful in treatment of cancers with defects in the
mitochondrial pathway.
Lysosomes, the underestimated suicide bags
Lysosomes have until recently been considered as
‘suicide bags’ that through the release of unspecific
enzymes cause autolysis and damage neighbouring cells
during uncontrolled tissue damage. However, accumu-
lating data now show that lysosomes also function as
death signal integrators in many controlled death
paradigms (Brunk et al., 2001; Ferri and Kroemer,
2001; Turk et al., 2002; Ja
¨
a
¨
ttela
¨
and Tschopp, 2003)
(Figure 2). Lysosomal proteases, cathepsins, translocate
from the lysosomal lumen to the cytosol in response to a
wide variety of death stimuli such as TNF and TRAIL
(Foghsgaard et al., 2001), Fas (Brunk and Svensson,
1999), p53 activation (Yuan et al., 2002), retinoids
(Zang et al., 2001), growth factor starvation and
oxidative stress (Brunk and Svensson, 1999), B-cell
receptor activation (van Eijk and de Groot, 1999), T-cell
activation and staurosporine (Bidere et al., 2003). Once
released to the cytosol, cathepsins, especially cysteine
cathepsins B and L and aspartyl cathepsin D, may
trigger MMP followed by caspase- or AIF-mediated
apoptosis (Roberg et al., 2002; Bidere et al., 2003; Boya
et al., 2003; Cirman et al., 2003). Akin to caspase-8,
lysosomal cysteine cathepsins can cleave and activate
Bid (Cirman et al., 2003). Thus, Bid may function as one
of the links that connects lysosomal leakage to MMP.
Lysosomal leakage may also trigger mitochondrial
dysfunction in a Bid- and Bcl-2-independent manner
Figure 2 Lysosomal control of PCD. Lysosomal leakage is
induced by the indicated stimuli by an undefined mechanism.
Effector molecules released from the lysosome include cathepsins,
ROS and H
þ
. Cathepsins released from the lysosomes function in
the apoptosis pathway by causing mitochondrial membrane
permeabilization via cleavage of Bid or via activation of PLA2
and the following increase in arachidonic acid (AA) level.
Cathepsins may also directly cause caspase-independent chromatin
condensation. Acidification of the cytosol can stabilize BNIP3 that
triggers autophagy or activate
L-DNase II that can cause partial
chromatinolysis in a caspase-independent manner. ROS may be
involved in mitochondrial damage and necrosis-like PCD
Cell death pathways as regulators of tumour initiation and progression
MJa¨a¨ttela¨
2750
Oncogene
(Boya et al., 2003). Second messengers involved in such
pathways may include ROS and arachidonic acid (Boya
et al., 2003; Foghsgaard et al., 2002). Cathepsins can
also mediate MMP- and caspase-independent PCD with
apoptosis-like morphology (Vancompernolle et al.,
1998; Foghsgaard et al., 2001). As the latter pathway
can circumvent most of the known resistance mechan-
isms occurring in tumour cells, lysosomal membrane
permeabilization appears as a promising target for
cancer therapy.
Lysosomal leakage leads inevitably to intracellular
acidification, whose role in lysosomal PCD has been
only poorly studied. The acidification of the cytosol has,
however, many consequences that may contribute to the
lysosomal PCD. Firstly, acidification has been shown to
trigger cJun N-terminal kinase (JNK)-mediated caspase-
and Bcl-2-independent necrosis-like PCD in bladder
cancer cells (Zanke et al., 1998). Secondly, acidification
can induce a post-translational modification of an
antiapoptotic serine protease inhibitor, leucocyte elas-
tase inhibitor, that converts it into
L-DNase II that in
turn can mediate nuclear changes typical of apoptosis-
like PCD (Altairac et al., 2003). And finally, autophagy-
inducing Bcl-2 family member BNIP3 can be stabilized
by intracellular acidification (Kubasiak et al., 2002).
Whether these mechanisms are involved in lysosomal
death pathways needs to be experimentally tested, but
they may explain some of the varying morphologies seen
following lysosomal leakage.
Lysosomes are also crucial in the autophagic degen-
eration (Lockshin and Zakeri, 2002; Reggiori and
Klionsky, 2002; Ogier-Denis and Codogno, 2003). Even
though autophagy can be initiated by mitochondrial
damage, it does also occur in the absence of MMP, and
most described forms of autophagic death are insensitive
to Bcl-2. During autophagy (or macroautophagy), a
portion of cytosol or cytosolic organelles are sequestered
by a structure known as autophagosome that is isolated
from the remaining cell contents by a double-membrane.
Subsequently, the autophagosome fuses with lysosomes
and the inner membrane as well as the sequestered
materials are digested by lysosomal enzymes. Morpho-
logically, autophagy is characterized by the appearance
of abundant vacuoles in the cytosol and it has been
described as a prominent form of PCD during develop-
ment (Lockshin and Zakeri, 2002). Of interest to cancer
research, autophagy-inducing stimuli include starvation,
hypoxia, radiation, antioestrogens as well as cytokines
such as TNF and interferon g (Ogier-Denis and
Codogno, 2003). Moreover, emerging data suggest that
autophagy may play a central role in the control of
tumorigenesis (see below).
Nucleus as a death sensor
The best-defined nuclear death pathway leading to
MMP and apoptosis is that triggered by DNA
damage-induced activation of p53 (Vousden and Lu,
2002). Bax and ‘BH3-only’ proteins Noxa and PUMA
are direct target genes of p53 and gene knockout studies
have revealed that they are required for efficient DNA
damage-induced MMP and apoptosis (Villunger et al.,
2003). Additionally, ROS may mediate p53-induced
MMP via the upregulation of genes involved in ROS
generation. Recent data also suggest that p53 may serve
as a link to the stress-induced direct activation of
caspase-2. p53-induced death domain protein (PIDD),
RAIDD (RIP-associated ICE-like death domain pro-
tein) and caspase-2 form an apoptosome-like complex
capable of activating caspase-2 (J Tschopp, personal
communication). The activated caspase-2 can then act
as an initiator caspase and trigger MMP to amplify the
caspase activation (Lassus et al., 2002). Whether the
stress-induced caspase-2 can mimic DR-activated cas-
pase-8 and engage a lethal caspase activation cascade
independent of MMP is as yet unclear.
Studies employing various pharmacological protease
inhibitors suggest that also noncaspase proteases with
activities resembling cysteine cathepsins (inhibited by z-
FA-fmk) and possibly other cysteine or serine proteases
(inhibited by TLCK) participate in p53-mediated PCD
(Lotem and Sachs, 1996). This may be explained by the
recent data showing that p53 provokes an early
lysosomal leakage that precedes MMP (Yuan et al.,
2002). Inhibitors of lysosomal cysteine proteases block
MMP in this model, suggesting that lysosomal leakage is
an early event that is necessary for the p53-induced
PCD. Additionally, DNA damage induced by g irradia-
tion, etoposide or adriamycin upregulates lysosomal
aspartyl protease cathepsin D in a p53-dependent
manner, and pepstatin A, a cathepsin D inhibitor,
partially suppresses p53-dependent death of lymphoid
cells (Wu et al., 1998). p53-induced lysosomal leakage is
not attenuated by an antioxidant implicating that ROS
are not involved in this process, whereas the ability of
caspase-2 and proapoptotic Bcl-2 family members to
destabilize lysosomes remains to be elucidated.
In spite of their importance in respect to the treatment
of human cancer, p53-independent apoptotic responses
to DNA damage are poorly studied. Possible pathways
involve activation of other tumour suppressor proteins;
for example, interferon regulatory factor 1 (Tamura
et al., 1995), breast cancer susceptibility gene BRCA1
(Harkin et al., 1999), p73 (Gong et al., 1999) and
promyelocytic leukaemia protein (PML) (Yang et al.,
2002). In addition to caspase-mediated apoptosis, PML
can trigger a caspase-independent death programme
displaying cytoplasmic apoptotic features, which are
even enhanced in the presence of pan-caspase inhibitors
(Quignon et al., 1998; Wang et al., 1998). Recent results
show also that DNA-damaging agents can trigger
lysosomal leakage in cancer cells lacking functional
p53 (J Nylandsted and MJ, unpublished).
ER as a sensor of stress
ER is an important sensor of cellular stress and it may
initiate PCD by at least two distinct mechanisms, the
unfolded protein response and release of Ca
2 þ
(Ferri
and Kroemer, 2001). Both events lead eventually to
MMP and may thus activate the classical apoptosis
pathway as well as the other death pathways triggered
Cell death pathways as regulators of tumour initiation and progression
MJa¨a¨ttela¨
2751
Oncogene
by MMP. Stimuli that induce an increase in the
intracellular free calcium [Ca
2 þ
]
i
may cause direct
calcium-mediated mitochondrial and/or lysosomal da-
mage or activate Bax via calpain-mediated cleavage
(Mattson et al., 2000; Choi et al., 2001; Leist and
Ja
¨
a
¨
ttela
¨
, 2001). Calpains may also convert Bcl-X
L
to a
proapoptotic MMP-inducing protein (Nakagawa and
Yuan, 2000).
Calpains are cysteine proteases that reside in the
cytosol in an inactive form, and their activation requires
an elevation in [Ca
2 þ
]
i
(Wang, 2000). Calpain activity is
controlled by calpastatin, a natural inhibitor that can be
inactivated by calpain- or caspase-mediated cleavage.
Calpains are activated by various stimuli (e.g. irradia-
tion, etoposide, neurotoxins and ionophores) that
increase the [Ca
2 þ
]
i
and they can in turn participate in
PCD signalling upstream or downstream of caspases
(Leist and Ja
¨
a
¨
ttela
¨
, 2001). Furthermore, calpains
mediate apoptosis-like PCD even in the absence of
caspase activation. For example, EB1089, a vitamin D
analogue currently in phase III clinical trials for the
treatment of cancer (Mathiasen et al., 1999; Mathiasen
et al., 2002) as well as re-expression of a tumor
suppressor gene ARHI (Bao et al., 2002) induce
calpain-dependent apoptosis-like PCD in cancer cells
without triggering detectable caspase activation.
Cytoskeleton and cytosol
The cytoskeleton controls MMP by binding to ‘BH3-
only’ proteins Bim and Bmf and keeping them thereby
inactive (Coultas and Strasser, 2003). Changes in the
cytoskeleton by various cytoskeleton disrupting or
stabilizing drugs or by cell detachment can thus lead
to MMP and apoptosis induced by the release of Bim
and Bmf. On the contrary, intact cytoskeleton is needed
for autophagy to occur (Ogier-Denis and Codogno, 2003).
Tumour cells are often challenged by hypoxia and
acidosis. Hypoxia-induced factor 1a (HIF-1a) upregu-
lates the autophagy-inducing Bcl-2 protein BNIP3
(Kothari et al., 2003), whose expression level is further
increased by acidosis-mediated stabilization (Kubasiak
et al., 2002). Accordingly, BNIP3 is expressed in hypoxic
regions of tumors but fails to induce cell death
presumably due to survival signalling via tyrosine kinase
receptors (Kothari et al., 2003). Notably, HIF-a also
upregulates many genes involved in apoptosis resistance,
metabolic adaptation, angiogenesis and metastasis
making the net effect of hypoxia most likely tumour
promoting (Semenza, 2003). Furthermore, HIF-1a is
overexpressed in human cancers not only due to hypoxia
but also as a result of increased synthesis induced by
oncogenes or decreased degradation due to loss of von
Hippel–Lindau tumour suppressor protein.
Taken together, all cellular compartments appear to
be equipped with sensors that can detect stressful
changes in the environment and weapons to set off
PCD providing a cell with multiple ways to get rid of
damaged and potentially dangerous cells. Thus, tumour
cells have to get rid of many enemies on their way
through transformation and tumour progression.
Promotion and inhibition of PCD during tumourigenesis
The intensive apoptosis research during the last 15 years
has convincingly demonstrated that the development of
aggressive tumours depends on numerous defects in
apoptosis signalling (Hanahan and Weinberg, 2000).
The requirement for apoptosis resistance is partially
explained by the nature of cancer-associated growth-
promoting signals themselves. Uncontrolled prolifera-
tive responses induced by enhanced activity of onco-
proteins like c-Myc, or the inactivation of tumour
suppressor proteins like retinoblastoma protein, result
in a deregulated cell cycle and sustained activation of
E2F-1, which can trigger caspase activation and
accelerated apoptosis (Evan et al., 1992; Wu and Levine,
1994) (Table 1). Even though growth promotion and
survival functions of oncoproteins of Ras and Src
families are prominent in tumours, they also carry the
ability to trigger cellular senescence and apoptosis and
may thus suppress transformation (Yao and Scott, 1993;
Frisch and Francis, 1994; Tanaka et al., 1994; Chen
et al., 1997). Many of the stresses that a cell must
encounter on its way to form an aggressive tumour
(hypoxia, starvation, cytotoxic lymphocytes, DNA
damage as well as detachment from the extracellular
matrix and neighbouring cells) further add to the
selective pressure to suppress apoptosis. To obtain net
growth, growth-stimulating signals and environmental
stresses must thus be coupled to reduced apoptosis
potential. In line with this, human tumours harbour
several inactivating mutations in proapoptotic genes
and/or show increased expression or activity of anti-
apoptotic proteins (Table 1). For example, mutations in
proapoptotic p53 tumour suppressor gene or changes in
the expression levels of its regulators are the most
frequent genetic abnormalities in human cancers (Vous-
den and Lu, 2002). Other common antiapoptotic
alterations in cancer cells include tilted balance in the
expression levels of pro- and antiapoptotic Bcl-2 family
members towards the favour of the latter; growth factor-
independent activation of receptor tyrosine kinases; lack
of PTEN, the negative regulator of phosphatidylinositol
3-kinase (PI3K) and protein kinase B (PKB, also called
Akt); activating mutations of PKB; inactivation of
Apaf-1 in therapy-resistant melanomas; overexpression
of survivin or other inhibitor of apoptosis protein family
members; and increased levels of anti-apoptotic heat-
shock proteins (Evan and Vousden, 2001; Reed, 2001;
Simpson and Parsons, 2001; Mathiasen and Ja
¨
a
¨
ttela
¨
,
2002; Coultas and Strasser, 2003; Downward, 2003).
Emerging evidence suggest that apoptosis is not the
only lethal challenge tumour cells encounter during
tumorigenesis. Transformation and tumour progression
are also coupled to alternative death pathways that have
to be suppressed for cancer to develop and evolve
(Table 1). For example, activating mutations of ras
proto-oncogenes, the most frequently mutated proto-
oncogenes in human tumors, trigger autophagy in
glioblastoma as well as colon and gastric cancer cells
(Chi et al., 1999; Pattingre et al., 2003). The ability of
Ras to cause autophagy may explain the autophagy
Cell death pathways as regulators of tumour initiation and progression
MJa¨a¨ttela¨
2752
Oncogene
observed in prostate cancer cells following activation of
ErbB2 and ErbB3 (receptor tyroisine kinases that
activate Ras pathway) (Tal-Or et al., 2003). Interest-
ingly, one of the signalling pathways regulated by Ras,
the Raf–Erk1/2 pathway, triggers autophagy (Pattingre
et al., 2003), whereas another downstream arm of Ras
signalling, PI3K–PKB pathway, inhibits it (Arico et al.,
2001). Additionally, the Ras-induced increase in the
expression and trafficking of lysosomal cathepsins may
contribute to Ras-induced autophagy (Roshy et al.,
2003). Cells that fail to escape autophagy by activation
of PI3K–PKB pathway, may do so by inactivating
mutations or depletions of genes encoding for Beclin-1,
Bin-1 or death-associated protein kinase (DAPK), all
proteins with tumour suppressor properties (Liang et al.,
1999; Elliott et al., 2000; Ogier-Denis and Codogno,
2003). Beclin-1 is a Bcl-2-interacting protein that has
structural similarity to the yeast autophagy gene apg6/
vps30. It is monoallelically deleted in 40–75% of
sporadic human breast cancers and ovarian cancers.
Bin1, on the other hand, interacts with c-Myc and
inhibits its transforming activity. It is frequently missing
or functionally inactivated in breast and prostate
cancers and in melanoma. DAPK is a calcium-regulated
serine/threonine kinase that in addition to activating the
p19ARF/p53-mediated apoptotic checkpoint, is an
important mediator of autophagy induced by starva-
tion, antioestrogens, interferon g and TNF (Raveh et al.,
2001; Inbal et al., 2002). Human tumour cell lines and
tumours of various origins have frequently lost DAPK
expression as a result of silencing by DNA methylation
(Raveh and Kimchi, 2001). Importantly, reintroduction
of Beclin-1, Bin-1 or an active form of DAPK into
tumour cells that either lack them or express them in low
levels triggers autophagic degeneration. Furthermore,
Bcl-2, the major oncogene controlling apoptosis, can
Table 1 Examples of cancer-related events that regulate PCD
Event Death pathway Effector Effect Reference
Activation of cancer-related proteins
Bcl-2 Apoptosis Caspases Inhibition Vaux et al. (1988)
Autophagy ND Inhibition Saeki et al. (2000)
Apoptosis-like Calcium Inhibition Lam et al. (1994)
c-Myc Apoptosis Caspases Sensitization Evan et al. (1992)
Apoptosis-like AIF Sensitization Liao and Dickson (2003)
E2F-1 Apoptosis Caspases Activation Wu and Levine (1994)
Src family Apoptosis Caspases Inhibition Frisch and Francis (1994)
Activation Yao and Scott (1993)
Apoptosis-like Cathepsins Sensitization Unpublished
Ras family Apoptosis Caspases Inhibition Frisch and Francis (1994)
Activation Tanaka et al. (1994)
Autophagy ND Activation Chi et al. (1999)
Apoptosis-like Cathepsins Sensitization Unpublished
NF-kB Apoptosis Caspases Inhibition Wang et al. (1996)
Apoptosis-like Cathepsins Inhibition Liu et al. (2003)
Raf–ERK1/2 Apoptosis Caspases Inhibition Kinoshita et al. (1997)
Autophagy ND Activation Pattingre et al. (2003)
PI3K–PKB Apoptosis Caspases Inhibition Kinoshita et al. (1997)
Autophagy ND Inhibition Arico et al. (2001)
Apoptosis-like Cathepsins Inhibition Madge et al. (2003)
Hsp70 Apoptosis Caspases Sensitization Liossis et al. (1997)
Apoptosis-like Cathepsins Inhibition Unpublished
Necrosis-like ROS Inhibition Creagh et al. (2000)
Cystatin A Apoptosis-like Cathepsins Inhibition Kuopio et al. (1998)
SV40 large T Autophagy ND Inhibition Elliott et al. (2000)
Cathepsins Apoptosis-like Cathepsins Sensitization Foghsgaard et al. (2001)
ErbB2/3 Autophagy ND Activation Tal-Or et al. (2003)
Loss of function of tumour suppressor-like proteins
p53 Apoptosis Caspases Inhibition Yonish-Rouach et al. (1991)
Apoptosis-like Cathepsins Inhibition Yuan et al. (2002)
RB Apoptosis Caspases Activation Wu and Levine (1994)
PTEN Apoptosis Caspases Inhibition Stambolic et al. (1998)
Autophagy ND Inhibition Arico et al. (2001)
Beclin-1 Autophagy ND Inhibition Liang et al. (1999)
PML Apoptosis Caspases Inhibition Wang et al. (1998)
Necrosis-like ND Inhibition Quignon et al. (1998)
Bin-1 Autophagy ND Inhibition Elliott et al. (2000)
DAPK Apoptosis Caspases Inhibition Raveh et al. (2001)
Apoptosis Caspases Activation Jin and Gallagher (2003)
Autophagy ND Inhibition Inbal et al. (2002)
ARHI Apoptosis-like Calpains Inhibition Bao et al. (2002)
JNK Apoptosis Caspases Inhibition Kennedy et al. (2003)
p27
KIP
Autophagy ND Inhibition Komata et al. (2003)
Cell death pathways as regulators of tumour initiation and progression
MJa¨a¨ttela¨
2753
Oncogene
also confer protection against autophagy (Saeki et al.,
2000). Downregulation of Bcl-2 in HL60 leukaemia cells
causes autophagy in a caspase- and MMP-independent
manner (Saeki et al., 2000). The cellular change
triggering autophagy in these cells is not defined, but it
could be related to the ability of Bcl-2 to control cellular
calcium homeostasis and thereby the activation of
DAPK (Lam et al., 1994). Finally, it should be noted
that autophagy does not necessarily always inhibit
tumorigenesis, but it may in fact also promote it via
its ability to protect against stresses associated with
tumour progression, that is hypoxia, starvation and
increased apoptosis (Ogier-Denis and Codogno, 2003).
Lysosomes and the control of their stability are
affected in many ways during the tumorigenesis.
Transformation by either Ras or Src upregulates the
expression of cysteine cathepsins, increases their secre-
tion and sensitizes cells to TNF-induced cathepsin B-
dependent PCD (Roshy et al., 2003) (N Fehrenbacher
and MJ, unpublished). Accordingly, tumour progres-
sion is associated with several changes that confer
resistance to the lysosomal death pathway, that is,
activation of PI3K and NF-kB, increased expression of
cathepsin inhibitors and Hsp70, as well as loss of p53
(Table 1). Hsp70 chaperone protein is commonly
overexpressed in human, tumours and its high expression
is associated with therapy resistance and poor prognosis
(Ja
¨
a
¨
ttela
¨
, 1999). Even though Hsp70 is mainly known as
a potent survival protein, it sensitizes cells to apoptosis
induced by Fas, T-cell receptor and cytotoxic lympho-
cytes suggesting that its protective function rather
covers nonapoptotic PCD pathways (Liossis et al.,
1997; Multhoff et al., 1997). Accordingly, the depletion
of Hsp70 induces a massive caspase-independent PCD
in tumour cells of various origins and experimental
tumours in mice, indicating that transformation is
coupled to the activation of an alternative Hsp70-
controlled death pathway (Nylandsted et al., 2000;
Nylandsted et al., 2002; Frese et al., 2003). This pathway
is mediated by lysosomal cysteine cathepsins that are
released into the cytosol of Hsp70-depleted tumour cells
before any signs of cell death are evident (J Nylandsted
and MJ, unpublished). Interestingly, Hsp70 is localized
to the lysosomal membranes of cancer cells and Hsp70
positive lysosomes display increased resistance against
chemical and physical membrane destabilization.
Furthermore, Hsp70 expression in diverse cell types
effectively inhibits TNF- and anticancer drug-induced
cell death at the level of the lysosomal leakage. Thus,
Hsp70 may promote tumorigenesis by stabilizing the
lysosomal membrane. Also, the mechanism by which
NF-kB helps cells to escape the lysosomal death was
recently uncovered (Liu et al., 2003). NF-kB induces the
expression of Spi2A, a potent inhibitor of serine
proteases and lysosomal cysteine cathepsins. Also, other
inhibitors of noncaspase proteases may confer growth
advantage to metastatic tumours that frequently express
high levels of cathepsins and other proteases capable of
degrading extracellular matrix (Duffy, 1996). For
example, the expression of Cystatin A in breast cancer
correlates with an aggressive phenotype and adverse
outcome (Kuopio et al., 1998), and tissue inhibitor of
metalloproteinases-1 rescues mammary epithelial cells
from stromelysin-1-induced PCD in vivo (Alexander
et al., 1996).
Concluding remarks
Death signalling in tumour cells appears much more
complex than originally suggested by the simple caspase
activation model, and PCD can result in multiple
morphological end points brought about by caspase-
independent mechanisms or by signalling routes utiliz-
ing caspases in combination with other effectors.
Collectively, the data indicate that various forms of
PCD are triggered by processes involved in transforma-
tion and tumour progression. Indeed, elimination of
cells that have lost functions of important tumour
suppressor proteins or bear activated oncogenes by PCD
may present the primary means by which such mutant
cells are continually removed from the body. The
participation of alternative PCD pathways together
with apoptosis in this process imposes a great challenge
to the developing tumour and may explain the relative
rarity of cancer in respect to the huge number of cell
divisions and mutations during a human life.
References
Alexander CM, Howard EW, Bissell MJ and Werb Z. (1996).
J. Cell Biol., 135, 1669–1677.
Alonso M, Tamasdan C, Miller DC and Newcomb EW.
(2003). Mol. Cancer Ther., 2, 139–150.
Altairac S, Zeggai S, Perani P, Courtois Y and Torriglia A.
(2003). Cell Death Differ., 10, 548–557.
Arico S, Petiot A, Bauvy C, Dubbelhuis PF, Meijer AJ,
Codogno P and Ogier-Denis E. (2001). J. Biol. Chem., 276,
35243–35246.
Ashkenazi A. (2002). Nat. Rev. Cancer, 2, 420–430.
Bao JJ, Le XF, Wang RY, Yuan J, Wang L, Atkinson EN,
LaPushin R, Andreeff M, Fang B, Yu Y and Bast Jr RC.
(2002). Cancer Res., 62, 7264–7272.
Bidere N, Lorenzo HK, Carmona S, Laforge M, Harper F,
Dumont C and Senik A. (2003). J. Biol. Chem., 278,
31401–31411.
Boya P, Andreau K, Poncet D, Zamzami N, Perfettini JL,
Metivier D, Ojcius DM, Ja
¨
a
¨
ttela
¨
M and Kroemer G. (2003).
J. Exp. Med., 197, 1323–1334.
Brunk UT, Neuzil J and Eaton JW. (2001). Redox Rep., 6,
91–97.
Brunk UT and Svensson I. (1999). Redox Rep., 4, 3–11.
Bursch W, Ellinger A, Kienzl H, Torok L, Pandey S, Sikorska
M, Walker R and Hermann RS. (1996). Carcinogenesis, 17,
1595–1607.
Chan HT, Hughes D, French RR, Tutt AL, Walshe CA,
Teeling JL, Glennie MJ and Cragg MS. (2003). Cancer Res.,
63, 5480–5489.
Charette SJ, Lambert H and Landry J. (2001). J. Biol. Chem.,
276, 36071–36074.
Chen G, Shu J and Stacey DW. (1997). Oncogene, 15,
1643–1651.
Cell death pathways as regulators of tumour initiation and progression
MJa¨a¨ttela¨
2754
Oncogene
Chi S, Kitanaka C, Noguchi K, Mochizuki T, Nagashima Y,
Shirouzu M, Fujita H, Yoshida M, Chen W, Asai A,
Himeno M, Yokoyama S and Kuchino Y. (1999). Oncogene,
18, 2281–2290.
Choi WS, Lee EH, Chung CW, Jung YK, Jin BK, Kim SU,
Oh TH, Saido TC and Oh YJ. (2001). J. Neurochem., 77,
1531–1541.
Cirman T, Oresic K, Droga Mazovec G, Turk V, Reed JC,
Myers RM, Salvesen GS and Turk B. (2003). J. Biol. Chem.,
,published online October 27.
Coultas L and Strasser A. (2003). Semin. Cancer Biol., 13,
115–123.
Creagh EM, Carmody RJ and Cotter TG. (2000). Exp. Cell
Res., 257, 58–66.
Downward J. (2003). Nat. Rev. Cancer, 3, 11–22.
Du C, Fang M, Li Y, Li L and Wang X. (2000). Cell, 102,
33–42.
Duffy MJ. (1996). Clin. Cancer Res., 2, 613–618.
Elliott K, Ge K, Du W and Prendergast GC. (2000). Oncogene,
19, 4669–4684.
Evan GI and Vousden KH. (2001). Nature, 411, 342–348.
Evan GI, Wyllie AH, Gilbert CS, Littlewood TD, Land H,
Brooks M, Waters CM, Penn LZ and Hancock DC. (1992).
Cell, 69, 119–128.
Ferri KR and Kroemer G. (2001). Nat. Cell Biol., 3,
E255–E263.
Foghsgaard L, Lademann U, Wissing D, Poulsen B and
Ja
¨
a
¨
ttela
¨
M. (2002). J. Biol. Chem., 277, 39499–39506.
Foghsgaard L, Wissing D, Mauch D, Lademann U, Bastholm
L, Boes M, Elling F, Leist M and Ja
¨
a
¨
ttela
¨
M. (2001). J. Cell
Biol., 153, 999–1009.
Forcet C, Ye X, Granger L, Corset V, Shin H, Bredesen DE
and Mehlen P. (2001). Proc. Natl. Acad. Sci. USA, 98,
3416–3421.
Frese S, Schaper M, Kuster JR, Miescher D, Ja
¨
a
¨
ttela
¨
M,
Buehler T and Schmid RA. (2003). J. Thorac. Cardiovasc.
Surg., 126, 748–754.
Frisch SM and Francis H. (1994). J. Cell Biol., 124,
619–626.
Gong JG, Costanzo A, Yang HQ, Melino G, Kaelin Jr WG,
Levrero M and Wang JY. (1999). Nature, 399, 806–809.
Hanahan D and Weinberg RA. (2000). Cell, 100, 57–70.
Harkin DP, Bean JM, Miklos D, Song YH, Truong VB,
Englert C, Christians FC, Ellisen LW, Maheswaran S,
Oliner JD and Haber DA. (1999). Cell, 97, 575–586.
Holler N, Zaru R, Micheau O, Thome M, Attinger A, Valitutti
S, Bodmer JL, Schneider P, Seed B and Tschopp J. (2000).
Nat. Immunol., 1, 489–495.
Inbal B, Bialik S, Sabanay I, Shani G and Kimchi A. (2002).
J. Cell Biol., 157, 455–468.
Ja
¨
a
¨
ttela
¨
M. (1999). Exp. Cell Res., 248, 30–43.
Ja
¨
a
¨
ttela
¨
M and Tschopp J. (2003). Nat. Immunol., 4,
416–423.
Jin Y and Gallagher PJ. (2003). J. Biol. Chem., published
online October 6.
Joseph B, Marchetti P, Formstecher P, Kroemer G, Lewen-
sohn R and Zhivotovsky B. (2002). Oncogene, 21, 65–77.
Kaufmann SH and Hengartner MO. (2001). Trends Cell Biol.,
11, 526–534.
Kelliher MA, Grimm S, Ishida Y, Kuo F, Stanger BZ and
Leder P. (1998). Immunity, 8, 297–303.
Kennedy NJ, Sluss HK, Jones SN, Bar-Sagi D, Flavell RA and
Davis RJ. (2003). Genes Dev., 17, 629–637.
Kerr JFR, Wyllie AH and Currie AR. (1972). Br. J. Cancer,
26, 239–257.
Khwaja A and Tatton L. (1999). J. Biol. Chem., 274, 36817–
36823.
Kinoshita T, Shirouzu M, Kamiya A, Hashimoto K,
Yokoyama S and Miyajima A. (1997). Oncogene, 15,
619–627.
Komata T, Kanzawa T, Takeuchi H, Germano IM, Schreiber
M, Kondo Y and Kondo S. (2003). Br. J. Cancer, 88, 1277–
1280.
Kothari S, Cizeau J, McMillan-Ward E, Israels SJ, Bailes M,
Ens K, Kirshenbaum LA and Gibson SB. (2003). Oncogene,
22, 4734–4744.
Kubasiak LA, Hernandez OM, Bishopric NH and Webster
KA. (2002). Proc. Natl. Acad. Sci. USA, 99, 12825–12830.
Kuopio T, Kankaanranta A, Jalava P, Kronqvist P, Kotkan-
salo T, Weber E and Collan Y. (1998). Cancer Res., 58,
432–436.
Lam M, Dubyak G, Chen L, Nun
˜
ez G, Miesfeld RL and
Distelhorst CW. (1994). Proc. Natl. Acad. Sci. USA, 91,
6569–6573.
Lamy L, Ticchioni M, Rouquette-Jazdanian AK, Samson M,
Deckert M, Greenberg AH and Bernard A. (2003). J. Biol.
Chem., 278, 23915–23921.
Lassus P, Opitz-Araya X and Lazebnik Y. (2002). Science,
297, 1352–1354.
Leist M and Ja
¨
a
¨
ttela
¨
M. (2001). Nat. Rev. Mol. Cell Biol., 2,
589–598.
Li LY, Luo X and Wang X. (2001). Nature, 412, 95–99.
Li P, Nijhawan D, Budihardjo I, Srinivasula SM, Ahmad M,
Alnemeri ES and Wang X. (1997). Cell, 91, 479–489.
Liang XH, Jackson S, Seaman M, Brown K, Kempkes B,
Hibshoosh H and Levine B. (1999). Nature, 402, 672–676.
Liao DJ and Dickson RB. (2003). Lab. Invest., 83, 1437–1449.
Liossis SN, Ding XZ, Kiang JG and Tsokos GC. (1997).
J. Immunol., 158, 5668–5675.
Liu N, Raja SM, Zazzeroni F, Metkar SS, Shah R, Zhang M,
Wang Y, Bromme D, Russin WA, Lee JC, Peter ME,
Froelich CJ, Franzoso G and Ashton-Rickardt PG. (2003).
EMBO J., 22, 5313–5322.
Lockshin RA and Zakeri Z. (2002). Curr. Opin. Cell Biol., 14,
727–733.
Lotem J and Sachs L. (1996). Proc. Natl. Acad. Sci. USA, 93,
12507–12512.
Luo X, Budihardjo I, Zou H, Slaughter C and Wang X.
(1998). Cell, 94, 481–490.
Luschen S, Ussat S, Scherer G, Kabelitz D and Adam-Klages
S. (2000). J. Biol. Chem., 275, 24670–24678.
Madge LA, Li JH, Choi J and Pober JS. (2003). J. Biol. Chem.,
278, 21295–21306.
Mateo V, Brown EJ, Biron G, Rubio M, Fischer A, Deist FL
and Sarfati M. (2002). Blood, 100, 2882–2890.
Mateo V, Lagneaux L, Bron D, Biron G, Armant M,
Delespesse G and Sarfati M. (1999). Nat. Med., 5,
1277–1284.
Mathiasen IS and Ja
¨
a
¨
ttela
¨
M. (2002). Trends Mol. Med., 8,
212–220.
Mathiasen IS, Lademann U and Ja
¨
a
¨
ttela
¨
M. (1999). Cancer
Res., 59, 4848–4856.
Mathiasen IS, Sergeev IN, Bastholm L, Elling F, Norman AW
and Ja
¨
a
¨
ttela
¨
M. (2002). J. Biol. Chem., 277, 30738–30745.
Mattson MP, LaFerla FM, Chan SL, Leissring MA,
Shepel PN and Geiger JD. (2000). Trends Neurosci., 23,
222–229.
Multhoff G, Botzler C, Jennen L, Schmidt J, Ellwart J and
Issels R. (1997). J. Immunol., 158, 4341–4350.
Nakagawa T and Yuan J. (2000). J. Cell Biol., 150,
887–894.
Nylandsted J, Rohde M, Brand K, Bastholm L, Elling F and
Ja
¨
a
¨
ttela
¨
M. (2000). Proc. Natl. Acad. Sci. USA, 97,
7871–7876.
Cell death pathways as regulators of tumour initiation and progression
MJa¨a¨ttela¨
2755
Oncogene
Nylandsted J, Wick W, Hirt UA, Brand K, Rohde M, Leist M,
Weller M and Ja
¨
a
¨
ttela
¨
M. (2002). Cancer Res., 62,
7139–7142.
Ogier-Denis E and Codogno P. (2003). Biochim. Biophys. Acta,
1603, 113–128.
Pattingre S, Bauvy C and Codogno P. (2003). J. Biol. Chem.,
278, 16667–16674.
Pettersen RD, Bernard G, Olafsen MK, Pourtein M and Lie
SO. (2001). J. Immunol., 166, 4931–4942.
Quignon F, De Bels F, Koken M, Feunteun J, Ameisen JC and
de The H. (1998). Nat. Genet., 20, 259–265.
Raveh T, Droguett G, Horwitz MS, DePinho RA and Kimchi
A. (2001). Nat. Cell Biol., 3, 1–7.
Raveh T and Kimchi A. (2001). Exp. Cell Res., 264, 185–192.
Reed JC. (2001). Trends Mol. Med., 7, 314–319.
Reggiori F and Klionsky DJ. (2002). Eukaryot. Cell, 1, 11–21.
Roberg K, Kagedal K and Ollinger K. (2002). Am. J. Pathol.,
161, 89–96.
Roshy S, Sloane BF and Moin K. (2003). Cancer Metast. Rev.,
22, 271–286.
Saeki K, Yuo A, Okuma E, Yazaki Y, Susin SA, Kroemer G
and Takaku F. (2000). Cell Death Differ., 7, 1263–1269.
Sakahira H, Enari M and Nagata S. (1998). Nature, 391,
96–99.
Scaffidi C, Fulda S, Srinivasan A, Friesen C, Li F, Tomaselli
KJ, Debatin KM, Krammer PH and Peter ME. (1998).
EMBO J., 17, 1675–1687.
Schulze-Osthoff K, Bakker AC, Vanhaesebroeck B, Beyaert R,
Jacob WA and Fiers W. (1992). J. Biol. Chem., 267,
5317–5323.
Scorrano L, Oakes SA, Opferman JT, Cheng EH, Sorcinelli
MD, Pozzan T and Korsmeyer SJ. (2003). Science, 300,
135–139.
Scorrano L, Penzo D, Petronilli V, Pagano F and Bernardi P.
(2001). J. Biol. Chem., 276, 12035–12040.
Semenza GL. (2003). Nat. Rev. Cancer, 3, 721–732.
Simpson L and Parsons R. (2001). Exp. Cell Res., 264, 29–41.
Sperandio S, de Belle I and Bredesen DE. (2000). Proc. Natl.
Acad. Sci. USA, 97, 14376–14381.
Stambolic V, Suzuki A, de la Pompa JL, Brothers GM,
Mirtsos C, Sasaki T, Ruland J, Penninger JM, Siderovski
DP and Mak TW. (1998). Cell, 95, 29–39.
Strasser A, O’Connor L and Dixit VM. (2000). Annu. Rev.
Biochem., 69, 217–245.
Susin SA, Lorenzo HK, Zamzami N, Marzo I, Snow BE,
Brothers GM, Mangion J, Jacotot E, Constantini P, Loeffer
M, Larochette N, Goodlett DR, Aebersold R, Siderovski
DP, Penninger JM and Kroemer G. (1999). Nature, 397,
441–446.
Suzuki Y, Imai Y, Nakayama H, Takahashi K, Takio K and
Takahashi R. (2001). Mol. Cell, 8, 613–621.
Tal-Or P, Di-Segni A, Lupowitz Z and Pinkas-Kramarski R.
(2003). Prostate, 55, 147–157.
Tamura T, Ishihara M, Lamphier MS, Tanaka N, Oishi I,
Aizawa S, Matsuyama T, Mak TW, Taki S and Taniguchi T.
(1995). Nature, 376, 596–599.
Tanaka N, Ishihara M, Kitagawa M, Harada H, Kimura T,
Matsuyama T, Lamphier MS, Aizawa S, Mak TW and
Taniguchi T. (1994). Cell, 77, 829–839.
Thome M and Tschopp J. (2001). Nat. Rev. Immunol., 1,
50–58.
Tolkovsky AM, Xue L, Fletcher GC and Borutaite V. (2002).
Biochimie, 84, 233–240.
Torriglia A, Perani P, Brossas JY, Chaudun E, Treton J,
Courtois Y and Counis MF. (1998). Mol. Cell. Biol., 18,
3612–3619.
Turk B, Stoka V, Rozman-Pungercar J, Cirman T, Droga-
Mazovec G, Oreic K and Turk V. (2002). Biol. Chem., 383,
1035–1044.
Vancompernolle K, Van Herreweghe F, Pynaert G, Van de
Craen M, De Vos K, Totty N, Sterling A, Fiers W,
Vandenabeele P and Grooten J. (1998). FEBS Lett., 438,
150–158.
Vande Velde C, Cizeau J, Dubik D, Alimonti J, Brown T,
Israels S, Hakem R and Greenberg AH. (2000). Mol. Cell.
Biol., 20, 5454–5468.
van Eijk M and de Groot C. (1999). J. Immunol., 163,
2478–2482.
Varfolomeev EE, Schuchmann M, Luria V, Chiannilkulchai
N, Beckmann JS, Mett IL, Rebrikov D, Brodianski VM,
Kemper OC, Kollet O, Lapidot T, Soffer D, Sobe T,
Avraham KB, Goncharov T, Holtmann H, Lonai P and
Wallach D. (1998). Immunity, 9, 267–276.
Vaux DL, Cory S and Adams JM. (1988). Nature, 335,
440–442.
Vercammen D, Beyaert R, Denecker G, Goossens V, Van Loo
G, Declercq W, Grooten J, Fiers W and Vandenabeele P.
(1998a). J. Exp. Med., 187, 1477–1485.
Vercammen D, Brouckaert G, Denecker G, Van de Craen M,
Declercq W, Fiers W and Vandenabeele P. (1998b). J. Exp.
Med., 188, 919–930.
Villunger A, Michalak EM, Coultas L, Mullauer F, Bock G,
Ausserlechner MJ, Adams JM and Strasser A. (2003).
Science, 302, 1036–1038.
Von Ahsen O, Waterhouse NJ, Kuwana T, Newmeyer DD
and Green DR. (2000). Cell Death Differ., 7, 1192–1199.
Vousden KH and Lu X. (2002). Nat. Rev. Cancer, 2,
594–604.
Wang C-Y, Mayo MW and Baldwin Jr AS. (1996). Science,
274, 784–787.
Wang KK. (2000). Trends Neurosci., 23, 20–26.
Wang X, Yang C, Chai J, Shi Y and Xue D. (2002). Science,
298, 1587–1592.
Wang ZG, Ruggero D, Ronchetti S, Zhong S, Gaboli M, Rivi
R and Pandolfi PP. (1998). Nat. Genet., 20, 266–272.
Wilson CA and Browning JL. (2002). Cell Death Differ., 9,
1321–1333.
Wu GS, Saftig P, Peters C and El-Deiry WS. (1998). Oncogene,
16, 2177–2183.
Wu X and Levine AJ. (1994). Proc. Natl. Acad. Sci. USA, 91,
3602–3606.
Yanagisawa H, Miyashita T, Nakano Y and Yamamoto D.
(2003). Cell Death Differ., 10, 798–807.
Yang S, Kuo C, Bisi JE and Kim MK. (2002). Nat. Cell Biol.,
4, 865–870.
Yao XR and Scott DW. (1993). Proc. Natl. Acad. Sci. USA,
90, 7946–7950.
Yeh W-C, de la Pompa JL, McCurrach ME, Shu H-B, Elia AJ,
Shahinian A, Ng M, Wakeman A, Khoo W, Mitchell K,
El-Deiry WS, Lowe SW, Goeddel DV and Mak TW. (1998).
Science, 279, 1954–1958.
Yonish-Rouach E, Resnitzky D, Lotem J, Sachs L, Kimchi A
and Oren M. (1991). Nature, 352, 345–347.
Yuan XM, Li W, Dalen H, Lotem J, Kama R, Sachs L and
Brunk UT. (2002). Proc. Natl. Acad. Sci. USA, 99,
6286–6291.
Zang Y, Beard RL, Chandraratna RA and Kang JX. (2001).
Cell Death Differ., 8, 477–485.
Zanke BW, Lee C, Arab S and Tannock IF. (1998). Cancer
Res., 58, 2801–2808.
Zhang J, Cado D, Chen A, Kabra NH and Winoto A. (1998).
Nature, 392, 296–300.
Cell death pathways as regulators of tumour initiation and progression
MJa¨a¨ttela¨
2756
Oncogene
... Lysosomes may play a dual role in cancer development [16,[39][40][41][42][43], because they can both exert an effect on invasive tumor growth and vascular development [16,44,45], as well as protect tumor cells from the effects of certain chemotherapeutic drugs, hence they may contribute to the development of drug resistance [16,44,45]. Lysosomes contribute to the maintenance of cancer cell proliferation, which may prevent oncogene-induced aging (OIS) [46]. ...
... However, it has been shown that cancer cells can inhibit both the apoptotic and lysosomal pathways in order to differentiate and proliferate [23]. An increased activity of phosphatidylinositol 3-kinase (PI3K) in cancer cells may affect many features of lysosomes, such as maturation rate, size, activity and stability [43,44]. On the other hand, the inhibition of the activity of phosphatidylinositol 3-kinase contributes to the change of the cell death pathway dependent on caspases to dependent on cathepsins [43]. ...
... An increased activity of phosphatidylinositol 3-kinase (PI3K) in cancer cells may affect many features of lysosomes, such as maturation rate, size, activity and stability [43,44]. On the other hand, the inhibition of the activity of phosphatidylinositol 3-kinase contributes to the change of the cell death pathway dependent on caspases to dependent on cathepsins [43]. Cancer cells can also protect themselves from lysosomal membrane damage by translocating a heat shock protein (Hsp 70) to their membrane. ...
Article
Full-text available
Lysosomes are organelles containing acidic hydrolases that are responsible for lysosomal degradation and the maintenance of cellular homeostasis. They play an important role in autophagy, as well as in various cell death pathways, such as lysosomal and apoptotic death. Various agents, including drugs, can induce lysosomal membrane permeability, resulting in the translocation of acidic hydrolases into the cytoplasm, which promotes lysosomal-mediated death. This type of death may be of great importance in anti-cancer therapy, as both cancer cells with disturbed pathways leading to apoptosis and drug-resistant cells can undergo it. Important compounds that damage the lysosomal membrane include lysosomotropic compounds, antihistamines, immunosuppressants, DNA-damaging drugs, chemotherapeutics, photosensitizers and various plant compounds. An interesting approach in the treatment of cancer and the search for ways to overcome the chemoresistance of cancer cells may also be combining lysosomotropic compounds with targeted modulators of autophagy to induce cell death. These compounds may be an alternative in oncological treatment, and lysosomes may become a promising therapeutic target for many diseases, including cancer. Understanding the functional relationships between autophagy and apoptosis and the possibilities of their regulation, both in relation to normal and cancer cells, can be used to develop new and more effective anticancer therapies.
... Evidence indicates that apoptosis is a homeostasis process to keep cell populations under control in some cancers, including glioblastoma (7) . Inhibition of apoptosis correlates with carcinogenesis and leads to drug resistance in glioblastoma (8)(9)(10)(11) . ...
... These detrimental substances set off a series of signaling pathways that ultimately lead to neurotoxicity and cell death. In order to combat the harmful effects of oxidative damage, the cellular redox balance is regulated by Nrf2, which acts as a protector (Jäättelä, 2004;Wang et al., 2019b). When Nrf2 is activated, individuals are shielded from injury caused by cerebral ischemia (Hahn et al., 2015). ...
Article
Reduced blood flow (hypoxia) to the brain is thought to be the main cause of strokes because it deprives the brain of oxygen and nutrients. An increasing amount of evidence indicates that the Centella-Asiatica (HA-CA) hydroalcoholic extract has a variety of pharmacological benefits, such as antioxidant activity, neuroprotection, anti-inflammatory qualities, and angiogenesis promotion. Intermittent fasting (IF) has neurological benefits such as anti-inflammatory properties, neuroprotective effects, and the ability to enhance neuroplasticity. The current study evaluates the combined effect of IF (for 1, 6, and 12 days) along with HA-CA (daily up to 12 days) in adult zebrafish subjected to hypoxia every 5 min for 12 days followed by behavioral (novel tank and open-field tank test), biochemical (SOD, GSH-Px, and LPO), inflammatory (IL-10, IL-1β, and TNF-α), mitochondrial enzyme activities (Complex-I, II, and IV), signaling molecules (AMPK, MAPK, GSK-3β, Nrf2), and imaging/staining (H&E, TTC, and TEM) analysis. Results show that sub-acute hypoxia promotes the behavioral alterations, and production of radical species and alters the oxidative stress status in brain tissues of zebrafish, along with mitochondrial dysfunction, neuroinflammation, and alteration of signaling molecules. Nevertheless, HA-CA along with IF significantly ameliorates these defects in adult zebrafish as compared to their effects alone. Further, imaging analysis significantly provided evidence of infarct damage along with neuronal and mitochondrial damage which was significantly ameliorated by IF and HA-CA. The use of IF and HA-CA has been proven to enhance the physiological effects of hypoxia in all dimensions.
... These drugs can enhance LMP or disrupt the activity of resident enzymes and protein complexes, such as v-ATPase and mTORC1 [48] (a list of inducers of lysosomal cell death can be found at, e.g., [49]). Mechanistically, an increase in cytosolic cathepsin activity triggers the mitochondrial membrane permeabilization through cleavage of Bid or via activation of phospholipase A2 and the consequent increase in araquidonic acid [50]. Furthermore, cathepsins can directly cause chromatin condensation [51], whereas the cytosol acidification can lead to L-DNAase II activation and chromatinolysis [52]. ...
Article
Full-text available
Simple Summary Moderate loco-regional hyperthermia (40–45 °C) is a therapeutic modality that can improve the effects of chemotherapy and radiotherapy and, in addition, improve our immune response against cancer. However, the effects of these combinations on many different cancers are modest. The use of much higher temperatures (>60 °C, thermal ablation) can cause severe damage to healthy tissues. In fact, our results show that cancer cells show extraordinary resistance to moderate hyperthermia. However, we found that the combination of hyperthermia (not higher than 52 °C) with low-strength electromagnetic fields acts synergistically, causing irreversible damage to different cancer cells. Moreover, these externally applied energies can be combined with chemotherapy and/or targeted therapies to achieve complete cancer eradication. In vivo, the energy causing focal hyperthermia can be distributed in multiple beams that can be concentrated in the tumor, thus avoiding damaging the healthy tissues that it passes through. Abstract At present, the applications and efficacy of non-ionizing radiations (NIR) in oncotherapy are limited. In terms of potential combinations, the use of biocompatible magnetic nanoparticles as heat mediators has been extensively investigated. Nevertheless, developing more efficient heat nanomediators that may exhibit high specific absorption rates is still an unsolved problem. Our aim was to investigate if externally applied magnetic fields and a heat-inducing NIR affect tumor cell viability. To this end, under in vitro conditions, different human cancer cells (A2058 melanoma, AsPC1 pancreas carcinoma, MDA-MB-231 breast carcinoma) were treated with the combination of electromagnetic fields (EMFs, using solenoids) and hyperthermia (HT, using a thermostated bath). The effect of NIR was also studied in combination with standard chemotherapy and targeted therapy. An experimental device combining EMFs and high-intensity focused ultrasounds (HIFU)-induced HT was tested in vivo. EMFs (25 µT, 4 h) or HT (52 °C, 40 min) showed a limited effect on cancer cell viability in vitro. However, their combination decreased viability to approximately 16%, 50%, and 21% of control values in A2058, AsPC1, and MDA-MB-231 cells, respectively. Increased lysosomal permeability, release of cathepsins into the cytosol, and mitochondria-dependent activation of cell death are the underlying mechanisms. Cancer cells could be completely eliminated by combining EMFs, HT, and standard chemotherapy or EMFs, HT, and anti-Hsp70-targeted therapy. As a proof of concept, in vivo experiments performed in AsPC1 xenografts showed that a combination of EMFs, HIFU-induced HT, standard chemotherapy, and a lysosomal permeabilizer induces a complete cancer regression.
... [46]). Mechanistically, an increase in cytosolic cathepsin activity triggers the mitochondrial membrane permeabilization through cleavage of Bid, or via activation of phospholipase A2 and the consequent increase in araquidonic acid [47]. Furthermore, cathepsins can directly cause chromatin condensation [48], whereas the cytosol acidification can also lead to L-DNAase II activation and chromatinolysis [49]. ...
Preprint
Full-text available
At present, the applications and efficacy of non-ionizing radiations (NIR) in oncotherapy are limited. In terms of potential combinations, the use of biocompatible magnetic nanoparticles as heat mediators has been extensively investigated. Nevertheless, the development of more efficient heat nanomediators that may exhibit high specific absorption rates still is an unsolved problem. Our aim was to investigate if externally applied magnetic fields and a heat-inducing NIR affect tumor cell viability. To this end, under in vitro conditions, different human cancer cells (A2058 melanoma, AsPC1 pancreas carcinoma, MDA-MB-231 breast carcinoma) were treated with the combination of electromagnetic fields (EMFs, using solenoids) and hyperthermia (HT, using a thermostated bath). The effect of NIR was also studied in combination with standard chemotherapy and targeted therapy. An experimental device combining EMFs and high intensity focused ultrasounds (HIFU)-induced HT was tested in vivo. EMFs (25 T, 4h) or HT (52C, 40 min) showed a limited effect on cancer cell viability in vitro. However, their combination decreased viability to approx. 16%, 50%, and 21% of controls values in A2058, AsPC1, and MDA-MB-231 cells, respectively. Increased lysosomal permeability, release of cathepsins into the cytosol and mitochondria-dependent activation of cell death are the underlying mechanisms. Cancer cells could be completely eliminated combining EMFs, HT and standard chemotherapy or EMFs, HT and anti-Hsp70-targeted therapy. As a proof of concept, in vivo experiments performed in AsPC1 xenografts showed that combination of EMFs, HIFU-induced HT, standard chemotherapy and a lysosomal permeabilizer induces a complete cancer regression.
... These drugs can enhance LMP or disrupt the activity of resident enzymes and protein complexes, such as v-ATPase and mTORC1 [48] (a list of inducers of lysosomal cell death can be found at, e.g., [49]). Mechanistically, an increase in cytosolic cathepsin activity triggers the mitochondrial membrane permeabilization through cleavage of Bid or via activation of phospholipase A2 and the consequent increase in araquidonic acid [50]. Furthermore, cathepsins can directly cause chromatin condensation [51], whereas the cytosol acidification can lead to L-DNAase II activation and chromatinolysis [52]. ...
Article
Full-text available
Introduction: Non-ionizing radiations (NIR), e.g. tumor-treating fields, high intensity focused ultrasounds (HIFU) or high frequency microwaves, have potential applications in Oncotherapy. Nevertheless, as today, their efficacy is limited. We investigated whether externally applied magnetic fields and non-ablative hyperthermia affect cancer cell viability and growth. Methods: Under in vitro conditions, human A2058 melanoma, AsPC1 pancreas carcinoma, and MDA-MB-231 breast carcinoma cells were treated with electromagnetic fields (EMFs, using solenoids) and/or hyperthermia (HT, using a thermostated water bath). The effect of both energies was also studied in combination with standard chemotherapy and targeted therapy. An experimental system combining EMFs and HIFU-induced HT was developed and tested in vivo. Results: EMFs (25 μT, 4h) or HT (52ºC, 40 min) quite moderately affect cancer cell viability. However, their combination decreased viability to approx. 16%, 50%, and 21% of controls values in A2058, AsPC1, and MDA-MB-231 cells, respectively). Increased lysosomal permeability, release of cathepsins into the cytosol and activation of the mitochondria-dependent cell death are the underlying mechanisms. Cancer cells were completely eliminated combining EMFs, HT and standard chemotherapy or EMFs, HT and anti-Hsp70-targeted therapy. In vivo experiments, using pancreatic cancer xenografts growing in nude mice, show that combination of EMFs, HIFU-induced HT and gemcitabine induces a drastic cancer regression (to approx. 3-4% of control values). Conclusion: Non-invasive NIR (EMFs and HIFU-induced HT) can complement current Oncotherapies. This strategy, associated to current chemotherapy or anti-Hsp70-targeted therapy, may facilitate the elimination of tumors that can be previously located using standard imaging techniques (CT scan and NMR). These results may help to overcome the limitations in the use of different NIR in Oncotherapy. Citation Format: Elena Obrador, Ali Jihad-Jebbar, Rosario Salvador-Palmer, Rafael López-Blanch, María Oriol-Caballo, María Paz Moreno-Murciano, Enrique Navarro, Rosa Cibrián, Jose M. Estrela. Electromagnetic fields and hyperthermia induce lysosomal permeabilization and death in different cancer cell models: in vitro and in vivo studies [abstract]. In: Proceedings of the American Association for Cancer Research Annual Meeting 2023; Part 1 (Regular and Invited Abstracts); 2023 Apr 14-19; Orlando, FL. Philadelphia (PA): AACR; Cancer Res 2023;83(7_Suppl):Abstract nr 2826.
... Additionally, LMP acts as an upstream signaling cascade for induction, MOMP, and the subsequent occurrence of apoptotic cell death (M. E. Guicciardi, M. Leist, & G. J. Gores, 2004;Jäättelä, 2004). Lipid mediator molecules, ROS (generated from metabolic pathways or excessive autophagosome accumulation), lysosomal hydrolases, cysteine proteases, and alleviated cytosolic Ca 2+ concentrations are efficient for inducing MOMP after LMP induction (Kroemer & Jäättelä, 2005). ...
Article
Recent developments in lysosome biology have transformed our view of lysosomes from static garbage disposals that can also act as suicide bags to decidedly dynamic multirole adaptive operators of cellular homeostasis. Lysosome‐governed signaling pathways, proteins, and transcription factors equilibrate the rate of catabolism and anabolism (autophagy to lysosomal biogenesis and metabolite pool maintenance) by sensing cellular metabolic status. Lysosomes also interact with other organelles by establishing contact sites through which they exchange cellular contents. Lysosomal function is critically assessed by lysosomal positioning and motility for cellular adaptation. In this setting, mechanistic target of rapamycin kinase (MTOR) is the chief architect of lysosomal signaling to control cellular homeostasis. Notably, lysosomes can orchestrate explicit cell death mechanisms, such as autophagic cell death and lysosomal membrane permeabilization‐associated regulated necrotic cell death, to maintain cellular homeostasis. These lines of evidence emphasize that the lysosomes serve as a central signaling hub for cellular homeostasis.
... The survival and death of cancer cells are regulated by several factors such as stress, damage, cell signaling, promotion of apoptosis, suppression of apoptosis, autophagy, and cell cycle progression. [16]. Suppression of apoptosis in a cell often leads to carcinogenesis. ...
Article
Purpose: To determine the effect of glycitin on PI3K/AKT signaling, migration, invasion, apoptosis, and cell cycle in A549 lung cancer cells.Methods: 3-[4,5-Dimethylthiazole-2-yl]-2,5-diphenyltetrazolium bromide (MTT) and clonogenic assays were used to determine the proliferation and colony generation potency of A549 cells, respectively, after treatment with glycitin (0 - 120 μM). Apoptosis in A549 cells was measured using DAPI and Annexin V/PI-FITC assays. Cell cycle arrest was assessed byflow cytometry, while the effect of glycitin on migration and invasion of A549 cells was determined by Transwell assay. The effect of glycitin on expressions of proteins associated with PI3K/AKT signaling in A549 cells was measured using western blotting.Results: Glycitin significantly inhibited the proliferation and colony generation potential of A549 cells (p < 0.05). The antiproliferative effects of glycitin on A549 cells were mediated through stimulation of apoptosis and cell cycle arrest at G0/G1-phase. The compound also distorted normal cellular morphology by causing membrane damage and nuclear fragmentation. The proportion of cells in the G0/G1-phase increased after glycitin treatment, when compared to the other two phases, demonstrating cell cycle arrest (p < 0.05). Glycitin suppressed the migration and invasion of A549 cells. However, Western blotting results showed that glycitin down-regulated the expressions of PI3K/AKT signaling proteins in A549 cells (p < 0.05).Conclusion: Glycitin produced significant anticancer effect on A549 cells via enhanced apoptosis, induction of cell cycle arrest, and inhibition of PI3K/AKT signalling. Moreover, it suppressed the migration and invasion of A549 cells. Therefore, glycitin is a potential lead molecule for the development of a therapeutic agent for invasive lung cancer.
... Although apoptotic pathways are widely described in literature [31][32][33][34][35] many aspects need to be considered in characterizing this type of cell death. Cationic substances are driven into mitochondria by the potential of mitochondrial membrane and this property can be used for intramitochondrial transport of drugs. ...
Article
Full-text available
Apoptosis is a widely controlled, programmed cell death, defects in which are the source of various diseases such as neurodegenerative diseases as well as cancer. The use of apoptosis in the therapy of various human diseases is of increasing interest, and the analysis of the factors involved in its regulation is valuable in designing specific carriers capable of targeting cell death. Highly efficient and precisely controlled delivery of genetic material by low-toxic carriers is one of the most important challenges of apoptosis-based gene therapy. In this work, we investigate the effect of the star polymer with 28 poly(N,N’-dimethylaminoethyl methacrylate) arms (STAR) on human cells, according to its concentration and structure. We show that star polymer cytotoxicity increases within its concentration and time of cells treatment. Except for cytotoxic effect, we observe morphological changes such as a shrinkage, loss of shape and begin to detach. We also prove DNA condensation after star polymer treatment, one of the most characteristic feature of apoptosis. The results indicate that the use of STAR triggers apoptosis in cancer cells compared to various normal cells, what makes these nanoparticles a promising drug in therapeutic strategy, which targets apoptosis. We demonstrate highlighting potential of star polymers as an innovative tool for anti-cancer therapy.
Article
Full-text available
The smart strategy of cancer cells to bypass the caspase-dependent apoptotic pathway has led to the discovery of novel anti-cancer approaches including the targeting of lysosomes. Recent discoveries observed that lysosomes perform far beyond just recycling of cellular waste, as these organelles are metabolically very active and mediate several signalling pathways to sense the cellular metabolic status. These organelles also play a significant role in mediating the immune system functions. Thus, direct or indirect lysosome-targeting with different drugs can be considered a novel therapeutic approach in different disease including cancer. Recently, some anticancer lysosomotropic drugs (eg, nortriptyline, siramesine, desipramine) and their nanoformulations have been engineered to specifically accumulate within these organelles. These drugs can enhance lysosome membrane permeabilization (LMP) or disrupt the activity of resident enzymes and protein complexes, like v-ATPase and mTORC1. Other anticancer drugs like doxorubicin, quinacrine, chloroquine and DQ661 have also been used which act through multi-target points. In addition, autophagy inhibitors, ferroptosis inducers and fluorescent probes have also been used as novel theranostic agents. Several lysosome-specific drug nanoformulations like mixed charge and peptide conjugated gold nanoparticles (AuNPs), Au-ZnO hybrid NPs, TPP-PEG-biotin NPs, octadecyl-rhodamine-B and cationic liposomes, etc. have been synthesized by diverse methods. These nanoformulations can target cathepsins, glucose-regulated protein 78, or other lysosome specific proteins in different cancers. The specific targeting of cancer cell lysosomes with drug nanoformulations is quite recent and faces tremendous challenges like toxicity concerns to normal tissues, which may be resolved in future research. The anticancer applications of these nanoformulations have led them up to various stages of clinical trials. Here in this review article, we present the recent updates about the lysosome ultrastructure, its cross-talk with other organelles, and the novel strategies of targeting this organelle in tumor cells as a recent innovative approach of cancer management.
Article
Full-text available
Cell death is achieved by two fundamentally different mechanisms: apoptosis and necrosis. Apoptosis is dependent on caspase activation, whereas the caspase-independent necrotic signaling pathway remains largely uncharacterized. We show here that Fas kills activated primary T cells efficiently in the absence of active caspases, which results in necrotic morphological changes and late mitochondrial damage but no cytochrome c release. This Fas ligand–induced caspase-independent death is absent in T cells that are deficient in either Fas-associated death domain (FADD) or receptor-interacting protein (RIP). RIP is also required for necrotic death induced by tumor necrosis factor (TNF) and TNF-related apoptosis-inducing ligand (TRAIL). In contrast to its role in nuclear factor κB activation, RIP requires its own kinase activity for death signaling. Thus, Fas, TRAIL and TNF receptors can initiate cell death by two alternative pathways, one relying on caspase-8 and the other dependent on the kinase RIP.
Article
Full-text available
Active cell death in hormone-dependent cells was studied using cultured human mammary carcinoma cells (MCF-7) treated with the anti-estrogens (AEs) tamoxifen (TAM), 4-hydroxy-tamoxifen (OH-TAM) or ICI 164 384 (10−8–10−5M) as a model. The following results were obtained. (i) In untreated MCF-7 cells a wave of replication occurred in the first 5 days of culture. All three AEs caused a dose-dependent inhibition of cell replication. (ii) TAM and OH-TAM at 10−5 M, but not ICI 164 384, caused lytic cell death (necrosis) within 24 h, which was not inhibited by estradiol (10−9–10−6M). (iii) Lower concentrations of TAM or OH-TAM (up to 10−6M) or ICI 164 384 induced a more gradual appearance of cell death beginning at day 3. This type of cell death was inhibited by estradiol (10−9 M), indicating its active nature. (iv) Nuclei showed two distinct patterns of alternation: (a) apoptosis-like condensation and fragmentation of chromatin to crescent masses abutting the nuclear envelope; (b) condensation of the chromatin to a single, pyknotic mass in the center of the nucleus, detached from the nuclear envelope. Quantitative histological evaluation revealed the predominance of pyknosis. (v) Biochemical DNA analysis revealed that only a relatively small amount of the total DNA was finally degraded into low molecular weight fragments (20 kb and less). (vi) Active cell death, with both apoptotic and pyknotic nuclear morphology, was associated with extensive formation of autophagic vacuoles (AV). 3-Methyladenine, a known inhibitor of AV formation, partially prevented cell death as detected by nuclear changes. (vii) ICI 164 384 was about 10 times more effective than TAM or OH-TAM at inhibiting DNA synthesis, but had equal potency in inducing active cell death. It is concluded that AEs have anti-proliferative and anti-survival effects on MCF-7 human mammary cancer cells in culture. These two effects are under separate control because they differ by kinetics, dose dependence and sensitivity to the various AEs. Active cell death in MCF-7 cells seems to be initiated by autophagy, in contrast to concepts of apoptosis, and thus corresponds to autophagic/lysosomal or type II death as previously defined. This may be important because of biochemical and molecular differences between these various subtypes of active cell death.
Article
Full-text available
We have identified two cell types, each using almost exclusively one of two different CD95 (APO-1/Fas) signaling pathways. In type I cells, caspase-8 was activated within seconds and caspase-3 within 30 min of receptor engagement, whereas in type II cells cleavage of both caspases was delayed for approximately 60 min. However, both type I and type II cells showed similar kinetics of CD95-mediated apoptosis and loss of mitochondrial transmembrane potential (DeltaPsim). Upon CD95 triggering, all mitochondrial apoptogenic activities were blocked by Bcl-2 or Bcl-xL overexpression in both cell types. However, in type II but not type I cells, overexpression of Bcl-2 or Bcl-xL blocked caspase-8 and caspase-3 activation as well as apoptosis. In type I cells, induction of apoptosis was accompanied by activation of large amounts of caspase-8 by the death-inducing signaling complex (DISC), whereas in type II cells DISC formation was strongly reduced and activation of caspase-8 and caspase-3 occurred following the loss of DeltaPsim. Overexpression of caspase-3 in the caspase-3-negative cell line MCF7-Fas, normally resistant to CD95-mediated apoptosis by overexpression of Bcl-xL, converted these cells into true type I cells in which apoptosis was no longer inhibited by Bcl-xL. In summary, in the presence of caspase-3 the amount of active caspase-8 generated at the DISC determines whether a mitochondria-independent apoptosis pathway is used (type I cells) or not (type II cells).
Article
The transcriptional activator interferon regulatory factor 1 (IRF-1) and its antagonistic repressor IRF-2 are regulators of the interferon (IFN) system and of cell growth. Here we report that embryonic fibroblasts (EFs) from mice with a null mutation in the IRF-1 gene (IRF-1−/− mice) can be transformed by expression of an activated c-Ha-ras oncogene. This property is not observed in EFs from wild-type or IRF-2−/− mice but is still observed in EFs from mice deficient in both genes. The transformed phenotype of ras-expressing IRF-1−/− EFs could be suppressed by the expression of the IRF-1 cDNA. Thus, IRF-1 functions as a tumor suppressor. Furthermore, expression of the c-Ha-ras oncogene causes wild-type but not IRF-1−/− EFs to undergo apoptosis when combined with a block to cell proliferation or treated by anticancer drugs or ionizing radiation. Hence, IRF-1 may be a critical determinant of oncogene-induced cell transformation or apoptosis.
Article
Two cysteine protease families, caspase and calpain, are known to participate in cell death. We investigated whether a stress-specific protease activation pathway exists, and to what extent Bcl-2 plays a role in preventing drug-induced protease activity and cell death in a dopaminergic neuronal cell line, MN9D. Staurosporine (STS) induced caspase-dependent apoptosis while a dopaminergic neurotoxin, MPP+ largely induced caspase-independent necrotic cell death as determined by morphological and biochemical criteria including cytochrome c release and fluorogenic caspase cleavage assay. At the late stage of both STS- and MPP+-induced cell death, Bax was cleaved into an 18-kDa fragment. This 18-kDa fragment appeared only in the mitochondria-enriched heavy membrane fraction of STS-treated cells, whereas it was detected exclusively in the cytosolic fraction of MPP+-treated cells. This proteolytic cleavage of Bax appeared to be mediated by calpain as determined by incubation with [35S]methionine-labelled Bax. Thus, cotreatment of cells with calpain inhibitor blocked both MPP+- and STS-induced Bax cleavage. Intriguingly, overexpression of baculovirus-derived inhibiting protein of caspase, p35 or cotreatment of cells with caspase inhibitor blocked STS- but not MPP+-induced Bax cleavage. This appears to indicate that calpain activation may be either dependent or independent of caspase activation within the same cells. However, cotreatment with calpain inhibitor rescued cells from MPP+-induced but not from STS-induced neuronal cell death. In these paradigms of dopaminergic cell death, overexpression of Bcl-2 prevented both STS- and MPP+-induced cell death and its associated cleavage of Bax. Thus, our results suggest that Bcl-2 may play a protective role by primarily blocking drug-induced caspase or calpain activity in dopaminergic neuronal cells.
Article
Research performed over the past decade has transformed apoptosis from a distinctive form of cell death known only by its characteristic morphology and genomic destruction to an increasingly well understood cellular disassembly pathway remarkable for its complex and multifaceted regulation. Here, we summarize current understanding of apoptotic events, note recent advances in this field and identify questions that might help guide research in the coming years.
Article
The death domain serine/threonine kinase RIP interacts with the death receptors Fas and tumor necrosis receptor 1 (TNFR1). In vitro, RIP stimulates apoptosis, SAPK/JNK, and NF-kappaB activation. To define the physiologic role(s) that RIP plays in regulating apoptosis in vivo, we introduced a rip null mutation in mice through homologous recombination. RIP-deficient mice appear normal at birth but fail to thrive, displaying extensive apoptosis in both the lymphoid and adipose tissue and dying at 1-3 days of age. In contrast to a normal thymic anti-Fas response, rip-/- cells are highly sensitive to TNFalpha-induced cell death. Sensitivity to TNFalpha-mediated cell death in rip-/- cells is accompanied by a failure to activate the transcription factor NF-kappaB.
Article
Programmed cell death, or apoptosis, is important in homeostasis of the immune system: for example, non-functional or autoreactive lymphocytes are eliminated through apoptosis. One member of the tumour necrosis factor receptor (TNFR) family, Fas (also known as CD95 or Apo-1), can trigger cell death and is essential for lymphocyte homeostasis. FADD/Mort1 is a Fas-associated protein that is thought to mediate apoptosis by recruiting the protease caspase-8. A dominant-negative mutant of FADD inhibits apoptosis initiated by Fas and other TNFR family members. Other proteins, notably Daxx, also bind Fas and presumably mediate a FADD-independent apoptotic pathway. Here we investigate the role of FADD in vivo by generating FADD-deficient mice. As homozygous mice die in utero, we generated FADD-/- embryonic stem cells and FADD-/- chimaeras in a background devoid of the recombination activating gene RAG-1, which activates rearrangement of the immunoglobulin and T-cell receptor genes. We found that thymocyte subpopulations were apparently normal in newborn chimaeras. Fas-induced apoptosis was completely blocked, indicating that there are no redundant Fas apoptotic pathways. As these mice age, their thymocytes decrease to an undetectable level, although peripheral T cells are present in all older FADD-/- chimaeras. Unexpectedly, activation-induced proliferation is impaired in these FADD-/- T cells, despite production of the cytokine interleukin (IL)-2. These results and the similarities between FADD-/- mice and mice lacking the beta-subunit of the IL-2 receptor suggest that there is an unexpected connection between cell proliferation and apoptosis.
Article
FADD (also known as Mort-1) is a signal transducer downstream of cell death receptor CD95 (also called Fas). CD95, tumor necrosis factor receptor type 1 (TNFR-1), and death receptor 3 (DR3) did not induce apoptosis in FADD-deficient embryonic fibroblasts, whereas DR4, oncogenes E1A and c-myc, and chemotherapeutic agent adriamycin did. Mice with a deletion in the FADD gene did not survive beyond day 11.5 of embryogenesis; these mice showed signs of cardiac failure and abdominal hemorrhage. Chimeric embryos showing a high contribution of FADD null mutant cells to the heart reproduce the phenotype of FADD-deficient mutants. Thus, not only death receptors, but also receptors that couple to developmental programs, may use FADD for signaling.