You are on page 1of 15

BBA - Molecular Cell Research 1867 (2020) 118688

Contents lists available at ScienceDirect

BBA - Molecular Cell Research


journal homepage: www.elsevier.com/locate/bbamcr

Review

Targeting apoptotic caspases in cancer T


a,b a,b,⁎
Ashley Boice , Lisa Bouchier-Hayes
a
Department of Pediatrics, Division of Hematology-Oncology and Department of Molecular and Cellular Biology, Baylor College of Medicine, Houston, TX 77030, USA
b
William T. Shearer Center for Human Immunobiology, Texas Children's Hospital, Houston, TX 77030, USA

A R T I C LE I N FO A B S T R A C T

Keywords: Members of the caspase family of proteases play essential roles in the initiation and execution of apoptosis. These
Caspase caspases are divided into two groups: the initiator caspases (caspase-2, -8, -9 and -10), which are the first to be
Cancer activated in response to a signal, and the executioner caspases (caspase-3, -6, and -7) that carry out the de-
Apoptosis molition phase of apoptosis. Many conventional cancer therapies induce apoptosis to remove the cancer cell by
IAPs
engaging these caspases indirectly. Newer therapeutic applications have been designed, including those that
Mitochondria
specifically activate individual caspases using gene therapy approaches and small molecules that repress natural
inhibitors of caspases already present in the cell. For such approaches to have maximal clinical efficacy,
emerging insights into non-apoptotic roles of these caspases need to be considered. This review will discuss the
roles of caspases as safeguards against cancer in the context of the advantages and potential limitations of
targeting apoptotic caspases for the treatment of cancer.

1. Introduction cause of death in the United States accounting for one in every 4 deaths
in the United States (www.cdc.gov/uscs.). One of the hallmarks of
Members of the caspase protease family play integral roles in the human cancer is the evasion of apoptosis [6]. Therefore, many existing
initiation and execution of apoptosis. Once activated, they cleave a therapies for cancer aim to overcome apoptosis evasion by targeting
number of structural and regulatory proteins to dismantle the cell from different caspase pathways. While most of these do so indirectly,
within [1]. These proteolytic events are responsible for the classical emerging insights into caspase functions and new therapeutic strategies
hallmarks of apoptosis including nuclear condensation, DNA fragmen- based on directly targeting caspases are beginning to show some pro-
tation, and plasma membrane blebbing [2]. Apoptosis plays crucial mise.
roles in the maintenance of cellular homeostasis and the removal of
damaged cells [3]. Not surprisingly, incorrect regulation of caspases 2. The caspase protease family
and apoptosis underlies the pathogenesis of many human diseases
[4,5]. The most notable of these is cancer. Cancer is the second leading The identification of the mammalian caspase family of proteases

Abbreviations: AIF, apoptosis inducing factor; ALL, acute lymphoblastic leukemia; AML, acute myeloid leukemia; APAF1, apoptotic peptidase activating factor 1;
ATM, ataxia-telangiectasia mutated; BAK, BCL2-antagonist killer 1; BAX, BCL2 associated X apoptosis regulator; BCL, pro-apoptotic B-cell lymphoma; BID, BH3-
interacting domain death agonist; BIR, baculovirus IAP repeat; CAD, caspase-activated DNase; CARD, caspase recruitment domain; c-FLIPL, cellular FLIP long; c-
FLIPS, cellular FLIP short; cGAS, cyclic GMP-AMP synthase; cIAP, cellular inhibitor of apoptosis protein; CICD, caspase-independent cell death; CID, chemical inducer
of dimerization; DD, death domain; DED, death effector domain; DIABLO, direct IAP binding protein with low pI; DISC, death inducing signaling complex; EndoG,
endonuclease G; FADD, fas-associated protein with death domain; FLIP, FLICE (FADD-like IL-1β-converting enzyme)-inhibitory protein; GAPDH, glyceraldehyde-3-
phosphate dehydrogenase; GvHD, graft-versus-host disease; HDAC, histone deacetylase; HER, human epidermal growth factor receptor; HOIL1, heme-oxidized IRP2
ubiquitin ligase 1; HOIP, HOIL1-interacting protein; HtrA2, high temperature requirement protein A2; IAP, inhibitor of apoptosis proteins; iCasp9, inducible version
of caspase-9; IKKγ, inhibitor of nuclear factor kappa-B kinase subunit gamma; iMOMP, incomplete MOMP; iPLA2, calcium-independent phospholipase A2; LUBAC,
linear ubiquitin assembly complex; MALT, mucosa-associated lymphoid tissue; MCL-1, myeloid leukemia cell differentiation protein; MDM2, murine double minute
2; MEF, mouse embryonic fibroblasts; MLKL, mixed lineage kinase domain like pseudokinase; MOMP, mitochondrial outer membrane permeabilization; MSC,
mesenchymal stromal cells; NEMO, NF-kappa-B essential modulator; NFκB, nuclear factor kappa-light-chain-enhancer of activated B cells; NIK, NFκB inducing
kinase; NSCLC, non-small-cell lung cancer; PARP, poly-ADP ribose polymerase; PIDD, p53-induced death domain protein; RasGAP, Ras GTPase; RhoGDI, Rho GDP-
disassociation inhibitor; RIPK1, receptor-interacting serine/threonine-protein kinase 1; ROCK1, Rho-associated coiled coil containing protein kinase 1; Sharpin,
SHANK Associated RH Domain Interactor; Smac, second mitochondria-derived activator of caspases; STING, stimulator of interferon genes

Corresponding author.
E-mail address: lxbouchi@txch.org (L. Bouchier-Hayes).

https://doi.org/10.1016/j.bbamcr.2020.118688
Received 16 September 2019; Received in revised form 20 January 2020; Accepted 15 February 2020
Available online 19 February 2020
0167-4889/ © 2020 Elsevier B.V. All rights reserved.
A. Boice and L. Bouchier-Hayes BBA - Molecular Cell Research 1867 (2020) 118688

Fig. 1. The human apoptotic caspases. A schematic representation of the domain organization of the human apoptotic caspases is shown. The caspases are divided
into the initiator caspases and the executioner caspases. Each caspase comprises a prodomain and a large and small catalytic subunit. Initiator caspases have long
prodomains containing the protein:protein interaction motifs, caspase recruitment domain (CARD) or death effector domain (DED), as indicated. The cleavage sites
for each caspase are depicted with red arrows.

Fig. 2. Structure of the caspase-3 heterotetramer. The crystal structure of the active caspase-2 heterodimer (left) and a ribbon diagram of the structure (right) are
shown. The large (p20) and small (p10) subunits of the heterodimer are depicted in blue and yellow respectively. Caspase-3 is shown bound to the inhibitor z-DEVD-
cmk (cyan). The catalytic histidine and cysteine are shown in red. The loops, L1 (magenta), L2 (black), LH (green), L3 (orange), and L4 (pale pink) make up the active
pocket. Caspase-3 is activated by cleavage of the intersubunit linker (L2). Cleavage of the intersubunit linker opens up the active pocket, exposing the active cysteine
(red star). For clarity, the loops, catalytic residues, and inhibitor are only highlighted in one half of the heterotetramer. The molecular models of the protein
structures were created using PyMOL v1.3.Edu with the full length caspase-3 crystal structure (PDB:2DKO) [38].

and their roles in apoptotic cell death was made possible by the dis- is generally required to form the stable active enzyme [10]. In order for
covery that the Caenorhabditis elegans gene Ced3 was required for a caspase to be activated, it must be in a dimer configuration that is
apoptosis in the nematode [7]. When the protein product of the Ced3 either preformed or induced [10]. The active enzyme is thus a het-
gene was cloned and sequenced, it was determined to be homologous to erotetramer of the catalytic domains comprised of two large and two
the mammalian cysteine protease caspase-1, then known as interleukin- small subunits (p20/p10)2 (Fig. 2) [11]. Each caspase has a specific
1β converting enzyme (ICE) [8]. Multiple additional mammalian pro- substrate specificity, but they all share the preference for an aspartate
teins were found to share similar homology with cell death protein 3 residue immediately prior to the cleavage site [12]. In rarer cases,
(CED-3) and this family of proteases became known as the caspases [9]. caspases can also cleave after glutamate and phosphoserine residues
Each caspase has in common an active site cysteine (‘C’) and pre- [13]. For the cysteine protease activity to occur, the caspase recognizes
ferentially cleaves its target substrates after an aspartate residue (‘asp- four residues (P4-P1, where P1 is the aspartate) on the N-terminal side
ase’) [9]. There are 14 known mammalian caspases and 12 known of a scissile bond (P1-P1′), resulting in cleavage of a peptide bond C-
human caspases. Caspases are zymogens comprised of three defined terminal to P1 [14].
domains: the prodomain, the large (~p20) catalytic domain, and small The mammalian caspases can be divided into three broad categories
catalytic domain (~p10) (Fig. 1). Cleavage between the three domains termed executioner caspases, initiator caspases, and inflammatory

2
A. Boice and L. Bouchier-Hayes BBA - Molecular Cell Research 1867 (2020) 118688

caspases. The executioner caspases include caspase-3 and caspase-7. responsible for the cleavage of the majority of apoptosis-associated
Caspase-6 is also in this group, but it has a less clearly defined role in substrates known [30,31].
apoptosis. Instead, it plays a unique role in neurodegenerative disease
[15]. The executioner caspases all have in common a short prodomain 3. Mechanisms of caspase activation
of 20–30 residues that lacks any defined motif or protein interaction
site. Executioner caspases are so-called because they bring about the The crystal structures of the catalytic domains of caspase-1, -2, -3,
execution phase of apoptosis by cleaving many different structural and -6, -7, -8, and -9 have been determined [11,32–38]. Many of the crystal
regulatory proteins that shut down the functions of the cell and con- structures of isolated caspases have demonstrated a stable structure as a
tribute to the morphological hallmarks of apoptosis [1]. The execu- homodimer of the catalytic domains representing the (p20/p10)2 het-
tioner caspases are activated when cleaved by an upstream initiator erotetramer. Each homodimer has an active site on either end of the
caspase or additional protease. Caspase-2, -8, -9, and -10 comprise the dimer. The homodimer is stabilized by hydrophobic interactions be-
initiator caspases. They each have a long prodomain that encodes a tween multiple beta-strands and one beta sheet of each monomer. The
protein:protein interaction motif used to facilitate dimerization of the active sites are made of 5 loops (L1–4 and LH) with L1 and L3 being the
catalytic domains. These motifs include the caspase recruitment domain more conserved of the loops and L2 and L4 more variant. L1 and L4
(CARD) found in caspase-2 and caspase-9, and the death effector do- make up either side of the substrate-binding pocket and L3 comprises
main (DED), found in the prodomains of caspase-8 and caspase-10 [16] the center of the pocket. L2 contributes the catalytic cysteine [39] and
(Fig. 1). These interaction domains allow the recruitment of the in- LH contributes a histidine [40] that together make up the catalytic dyad
itiator caspases to large multi-protein complexes that serve as activa- (Fig. 2). Executioner caspases are present in the cells as preformed di-
tion platforms for each caspase. Activation platforms are specific to mers that are activated by cleavage [10]. They are activated when the
each caspase and include the death inducing signaling complex (DISC) linker between the large and small subunit is cleaved by an upstream
for caspase-8 and caspase-10 [17], the apoptotic peptidase activating caspase or another protease such as granzyme B [41]. When the caspase
factor 1 (APAF1) apoptosome for caspase-9 [18], and the p53-induced is in its uncleaved form, the active site is occupied blocking the for-
death domain protein (PIDD) osome for caspase-2 [19]. Each complex mation of a functional substrate-binding site. The site is occupied by the
generally consists of a receptor protein that oligomerizes and recruits interdomain linker in caspase-3 and caspase-7 [35,42,43]. Cleavage of
either the caspase or an adaptor protein. For example, the CD95/Fas the linker forms L2 and L2′. L2′ stabilizes the active site of the neigh-
DISC is formed when oligomers of CD95/Fas recruit molecules of the boring dimer in the heterotetramer [42]. The cleavage and the ac-
adaptor protein Fas-associated protein with death domain (FADD). companying conformational change allows access to substrate binding
FADD in turn binds to caspase-8, a binding that is mediated by the DED pocket and stabilization of the active site (Fig. 2).
in FADD binding to the DED in caspase-8 [20]. Similarly, oligomer- Unlike the executioner caspases, the initiator caspases are activated
ization of APAF1 forms the apoptosome, providing the activation by dimerization rather than cleavage. The active site pocket is thus
platform for the recruitment of caspase-9 via the CARD present in created and exposed upon formation of the caspase dimer. Caspase-9
caspase-9 and in APAF1 [21]. Recruitment of caspase monomers to can be activated in the absence of cleavage [44], and dimeric caspase-2
their respective activation platform increases the local concentration of retains 20% activity when cleavage is blocked [22]. It has been shown
caspase molecules permitting dimerization. Dimerization is essential to in vitro that enforced dimerization of caspase-8 and caspase-9 resulted
initiate the activation of the caspase [22,23]. This mechanism is re- in enhanced activity. Blocking cleavage of caspase-8 impaired the sta-
ferred to as the induced proximity model [24]. A third group of cas- bility of the dimeric species [10]. Thus, dimerization of caspase-8 is
pases is known as the inflammatory caspase subgroup. These caspases required for activation, and cleavage serves to stabilize and further
consist of human caspases-1, -4, -5, and 12, murine caspase-11, and enhance activity of the enzyme. Similarly, cleavage of the interchain
bovine caspase-13. Each of these caspases has a long prodomain en- linker of capase-9 yielded a 5-fold increase in activity, while removal of
coding a CARD and, thus, they can be considered initiator caspases. the CARD resulted in a further 10-fold enhancement.
However, the inflammatory caspases are not generally involved in the The two main pathways of caspase activation that result in apop-
initiation of apoptotic cell death. Rather, they are defined by their tosis are termed the intrinsic pathway and the extrinsic pathway, so-
contribution to inflammation by regulating the proteolytic cleavage of called because they are generally engaged by stresses that come from
the immature cytokines, proIL-1β and proIL-18, or the pore forming outside the cells or within the cell respectively. The intrinsic pathway is
protein gasdermin D to induce an inflammatory form of cell death primarily activated by cellular stresses including DNA damage, meta-
called pyroptosis [25]. bolic stress, and endoplasmic reticulum stress. Many chemotherapeutic
In addition to their modes of activation, the separation into the drugs that are used to treat various cancers engage this pathway. These
initiator and executioner caspase groups is also based on their substrate stimuli converge on the mitochondria leading to mitochondrial outer
preferences. Caspase-3 and caspase-7 have similar preferences for membrane permeabilization (MOMP) and the release of cytochrome c
substrate sequences (DEXD), as determined using positional scanning from the mitochondrial intermembrane space to the cytosol (Fig. 2).
peptide libraries. Caspase-6, in contrast, appears to prefer branched Once in the cytosol, cytochrome c binds to the WD repeats of APAF1,
aliphatic residues in the P4 position. Caspase-8, caspase-9, caspase-10, inducing a conformational change that opens up the molecule. In the
and the inflammatory caspases all prefer substrates with an aliphatic presence of dATP or ATP, APAF1 then oligomerizes to assemble the
residue at P4 such as leucine or valine. The substrate specificity of APAF1 apoptosome [46,47]. In turn, the apoptosome recruits and ac-
caspase-2 is similar to those of common with caspase-3 and -7 [12,26]. tivates caspase-9, which continues the caspase cascade by cleaving the
For that reason, caspase-2 has sometimes been considered an execu- executioner caspases-3 and -7. To be able to cleave caspase-3, caspase-9
tioner caspase. However, it contains a long prodomain in its structure, needs to remain tethered to the apoptosome and caspase-9 auto-
which places it within the initiator caspase subgroup. It must be noted cleavage to the p35/p17 subunits induces dissociation from the apop-
that these substrate specificities are not always matched by the re- tosome [48]. This has been proposed to act as a molecular timer for
pertoire of natural substrates for each caspase [27]. For example, while caspase activation [49] rather than increasing the enzyme's activity as
caspase-3 and caspase-7 have similar peptide preferences they have few is seen with caspase-8.
substrates in common. These common substrates include poly-ADP ri- The extrinsic pathway generally refers to apoptosis induced by the
bose polymerase (PARP), Rho GDP-disassociation inhibitor (RhoGDI) engagement of death receptors by their cognate ligands at the cell
and Rho-associated coiled coil containing protein kinase 1 (ROCK1). membrane. The death receptors are all members of the tumor necrosis
Caspase-6 has a more limited substrate repertoire that includes lamin A factor receptor (TNFR) superfamily, but the death receptors have in
and ceramide phosphoethanolamine synthase [28,29]. Caspase-3 is common a death domain (DD) at their C-terminal. The DD is a protein

3
A. Boice and L. Bouchier-Hayes BBA - Molecular Cell Research 1867 (2020) 118688

Fig. 3. The mitochondrial pathway of apoptosis.


Diverse cellular stresses such as DNA damage, ER
stress, and metabolic stress activate the intrinsic
pathway leading to mitochondrial outer membrane
permeabilization (MOMP). MOMP allows for the
release of cytochrome c, second mitochondria-de-
rived activator of caspases/direct IAP binding pro-
tein with low pI (Smac/DIABLO), and Omi/high
temperature requirement protein A2 (HtrA2) from
the mitochondrial intermembrane space.
Cytochrome c binds to the WD repeats (red rec-
tangle) of apoptotic peptidase activating factor 1
(APAF1), inducing a conformational change that
opens up the molecule. This allows multiple mole-
cules of APAF1 to oligomerize and assemble into the
APAF1 apoptosome. The apoptosome recruits and
activates caspase-9, which then activates caspase-3
and caspase-7 leading to apoptosis. Smac/DIABLO
and Omi/HtrA2 bind to XIAP to prevent it from in-
hibiting active caspases.

interaction domain that shares structural features with the DED and from the intrinsic and extrinsic pathway. Unlike caspase-8 and caspase-
CARD. Death receptors include CD95, TNFR1, and the TNF-related 9, caspase-2 does not directly cleave the executioner caspases, but leads
apoptosis-inducing ligand (TRAIL) receptors (TRAIL-R1/DR4 and to their activation indirectly through the cleavage of the pro-apoptotic
TRAIL-R2/DR5). Each of these receptors is activated following multi- B-cell lymphoma 2 (BCL2) family protein, BH3-interacting domain
merization upon binding of their cognate ligand: CD95L/Fas ligand death agonist (BID), leading to MOMP and subsequent caspase-9 and
(FasL), TNF, and TRAIL respectively, which ultimately leads to activa- caspase-3 activation [55]. Caspase-8 can also cleave BID to induce
tion platform assembly and caspase-8 activation (Fig. 3). CD95 and MOMP, bridging the extrinsic and intrinsic apoptotic pathway [56].
TRAIL each assemble to form the DISC at the plasma membrane. In Hence, in certain cell types, caspase-8-induced apoptosis can be in-
contrast, the caspase-8 activation platform induced by the TNFR1 does hibited by BCL2 expression. Cells in which caspase-8 induced apoptosis
not initially assemble at the membrane. Upon engagement of TNFR1 by is not inhibited by anti-apoptotic BCL2-family proteins are termed type
TNF, the receptor rapidly recruits TRADD, a FADD-like homolog, re- I cells and cells that require MOMP for caspase-8-induced apoptosis are
ceptor-interacting serine/threonine-protein kinase 1 (RIPK1) and TNF called type II cells [57].
receptor associated factor 2 (TRAF2) at the plasma membrane [50].
Termed complex I, this complex induces nuclear factor kappa B (NFκB) 4. Mechanisms of caspase inhibition
activation. A second complex, complex IIa, is formed when TRADD/
RIPK1/TRAF2 dissociates away from TNFR1 and recruits FADD and There are a number of inhibitors of caspases that are expressed by
caspase-8 in the cytoplasm to induce apoptosis [50] (Fig. 3). If caspase- cells. While the function of these inhibitors is to limit uncontrolled
8 is absent or inactive, complex IIb is formed between RIPK1 and caspase activation and unnecessary cell death, high expression of cas-
RIPK3. RIPK3 induces phosphorylation of the pore forming protein, pase inhibitors in cancer cells is often associated with chemoresistance.
mixed lineage kinase domain like pseudokinase (MLKL), which forms The first group of inhibitors comprises the inhibitor of apoptosis pro-
pores in the plasma membrane to induce an alternative form of cell teins (IAPs), which includes X-linked IAP (XIAP), cIAP1, and cIAP2. The
death called necroptosis [51] (Fig. 3). IAPs are characterized by baculovirus IAP repeat (BIR) motifs con-
The adaptor molecule FADD recruits caspase-8 to the DISC or TNFR taining 1–3 Zn fingers and a RING finger domain on the carboxy-
complex II. Caspase-8 has two DEDs in its prodomain and the current terminal. Members of this group of proteins can inhibit caspases both
model for its activation shows that, when recruited to the DISC, these directly and indirectly. Human XIAP, cIAP1 and cIAP2 have all been
DEDs form filaments. Cryo-electron microscopy has revealed that the shown to bind to and inhibit the active forms of caspase-3, caspase-7,
caspase-8 DED filaments form ~20 nm wide structures [52]. The first and caspase-9 in vitro [58,59]. However, cIAP1 and cIAP2 bind caspases
DED of caspase-8 binds to FADD and the second DED in the same with much less efficiency because they lack important caspase inter-
caspase-8 molecule binds to the first DED of an additional caspase-8 acting residues that allow for tight binding and direct inhibition [60].
molecule through tandem linkage [53,54]. Two additional interaction Therefore, XIAP is the primary IAP to directly inhibit caspases in vivo.
interfaces of the two caspase-8 DED domains further amplify the DED The crystal structure of caspase-3 or caspase-7 in complex with XIAP
chains in two additional directions forming a helical filament assembly revealed that the BIR2 domain in the N-terminal portion of the XIAP
[52]. This provides a mechanism of amplifying the caspase-8 signal. protein binds to and blocks the substrate binding pocket of the caspase,
The pathway of caspase-2-induced apoptosis is somewhat distinct therefore obstructing recruitment of its substrates [61,62]. Caspase-9

4
A. Boice and L. Bouchier-Hayes BBA - Molecular Cell Research 1867 (2020) 118688

interacts with the BIR3 domain of XIAP [37]. XIAP is an E3 ligase. This primarily by binding to the DED of caspase-8 [77]. Certain studies have
provides a secondary means to inhibit caspase-induced apoptosis by shown that high levels of c-FLIPS and c-FLIPL result in shorter DED fi-
targeting active caspase-3 to the proteasome for degradation [63]. laments [77,82]. However, confocal imaging as well as mathematical
XIAP mediated inhibition of caspases is regulated by second mi- and structural modeling propose that c-FLIP inhibits caspase-8 by in-
tochondria-derived activator of caspases (Smac)/direct IAP binding creasing the proportion of caspase-8/FLIP heterodimers compared to
protein with low pI (DIABLO) and the related protein Omi/High tem- caspase-8 homodimers in the chain without affecting filament length
perature requirement protein A2 (HtrA2) [64–67]. Both of these pro- [52,54,83]. The reasons for these discrepancies are unclear but are
teins are released from the mitochondria along with cytochrome c at the likely due, in part, to the limitations of working with in vitro recon-
time of MOMP. Smac and Omi bind to the BIR2 or BIR3 domain of XIAP stituted complexes. When present at low concentrations or with high
thereby preventing its binding to the active caspases (Fig. 3) [67–69]. receptor stimulation, the formation of a c-FLIPL/caspase-8 heterodimer
The ability of Smac to modulate XIAP activity plays an important role in at the DISC promotes caspase-8 oligomers. Structural modeling suggests
the caspase-8 pathway. The difference between type I and type II cells is that c-FLIP isoforms may even stabilize the DISC by bridging neigh-
determined by the level of XIAP expression. In type I cells, XIAP ex- boring DISCs together [82]. Thus, in these conditions, c-FLIPL augments
pression is low or absent and therefore caspase-8-mediated activation of apoptosis [84,85].
caspase-3 is unimpeded and cell death is rapid. In contrast, in type II Caspase-10 is a close homolog of caspase-8 that also appears to
cells, XIAP expression is high. In these cells Smac release from mi- function as a negative regulator of the extrinsic pathway. Unlike, c-
tochondria is required to prevent XIAP from blocking caspase-8-medi- FLIPL, caspase-10 is catalytically active [86]. Caspase-10 has been
ated caspase-3 activation and apoptosis proceeds more slowly [70]. shown to inhibit CD95L-mediated death independently of c-FLIP by
cIAP1 and cIAP2 inhibit caspases in an indirect manner. cIAP1/2 binding to caspase-8 and preventing caspase-8 activation within the
block the formation of complex II, the caspase-8 activation complex DISC. Instead, recruitment of caspase-10 appears to have a pro-survival
induced by TNF. Following complex I formation, TRAF2 serves as a effect by promoting CD95L-mediated NFκB activation [87].
linker to recruit cIAP1 and cIAP2 to TNFR complex I [71] and, through Increased expression levels of each of these inhibitors have been
ubiquitination of components of the complex, cIAP1/2 activates the noted in a number of cancers. They also provide viable drug targets for
canonical NFκB pathway. Additionally, cIAP1/2 blocks complex II removing existing brakes on caspase activation and lowering the
formation by recruiting linear ubiquitin assembly complex (LUBAC) to threshold for apoptosis. Increased expression of c-FLIP or the specific c-
complex I. LUBAC is a complex that consists of the proteins heme- FLIP isoforms has been demonstrated in multiple cancers including
oxidized IRP2 ubiquitin ligase 1 (HOIL1), HOIL1-interacting protein lymphoma and non-small-cell lung cancer (NSCLC) [88–91] and its
(HOIP) and SHANK Associated RH Domain Interactor (SHARPIN). overexpression has been linked to poor prognosis and lower survival
LUBAC adds linear ubiquitin chains to inhibitor of nuclear factor kappa- rates [92–94]. Similarly, multiple lines of evidence show that IAP ex-
B kinase subunit gamma/NF-kappa-B essential modulator (IKKγ/ pression levels are an important determinant of chemoresistance in
NEMO), the regulatory component of the IKK complex [72]. This brings cancer. IAP overexpression has been noted in a number of human
the IKK complex into the TNFR1 complex, facilitating canonical NFκB cancers and is associated with increased drug resistance and poorer
activation resulting in the translocation of the NFκB p50/p65 com- outcomes [95,96].
plexes to the nucleus to induce transcription of multiple survival genes
including cIAP2. LUBAC also adds linear ubiquitin chains to other 5. Can caspases act as tumor suppressors?
components of the complex, stabilizing complex I and effectively in-
hibiting complex II formation (Fig. 4) [73]. Therefore, cIAP1/2 are key While many upstream regulators of caspases have been defined as
factors in the inhibition of caspase-8 activation through the TNF oncogenes or tumor suppressors [97], there have only been a few cases
pathway. cIAP1/2 also prevents non-canonical NFκB signaling by fa- where caspases themselves have been shown to function as tumor
cilitating degradation of NFκB inducing kinase (NIK) [74]. In the ab- suppressors (see below). In addition, there are not an overwhelming
sence of cIAP1/2, NIK phosphorylates IKKα inducing RelB/p52 com- number of published examples of deregulated caspase expression or
plexes. mutation profoundly impacting cancer incidence or progression. This
A second mechanism that cells have evolved to negatively regulate may be because of the essential nature of these proteins for develop-
caspases is by blocking dimerization of initiator caspases. FADD-like IL- ment. Genetic deletion of many caspases leads to some form of ab-
1Beta-converting enzyme-inhibitory protein (FLIP) is a caspase in- normality or lethality. Deletion of caspase-9 or its upstream regulator,
hibitor that has multiple isoforms. Cellular FLIP long (c-FLIPL) is a APAF1 results in perinatal lethality due to forebrain outgrowth and
catalytically inactive homolog of caspase-8. Cellular FLIP short (c- resistance to apoptotic inducers that trigger the intrinsic pathway
FLIPS) [75], and cellular FLIP raji (c-FLIPR [76]) only comprise the DED [98,99]. Caspase-3 knockouts generated on the murine 129 background
containing prodomain and lack the catalytic domains. c-FLIP can form showed a similar phenotype with exencephaly, perinatal lethality, and
heterodimers with caspase-8 [23]. Heterodimers formed between cas- delayed growth [100]. Surprisingly, thymocytes from the Casp3−/−
pase-8 and c-FLIPS are completely inactive, blocking caspase-8 activa- animals were not resistant to apoptosis in this model. When back-
tion [77]. In contrast, heterodimers between c-FLIPL and caspase-8 have crossed to B6 mice, the gross craniofacial abnormalities of Casp3−/−
catalytic activity. Heterodimerization between c-FLIPL and caspase-8 were ablated and the mice were born at expected Mendelian ratios
enables the stabilization of the active site without caspase-8 cleavage [101]. This indicates the existence of genetic modifiers in 129 mice that
and is energetically favoured when compared to the homodimerization causes loss of caspase-3 to induce exencephaly. Casp6−/− mice are
of caspase-8 [78]. The resulting c-FLIPL/caspase-8 complex has a more normal with only a slight resistance to anti-CD95-induced death [102].
limited substrate repertoire [79]. In the context of TNFR complex II, c- Knockout of caspase-7 on the B6 background was viable but combined
FLIPL/caspase-8 heterodimers result in the destabilization of the deletion of Casp3 and Casp7 resulted in lethality shortly after birth
RIPK1/RIPK3 complex (complex IIb), blocking RIPK3 signaling and [103]. Exencephaly was largely absent in these double knockout ani-
necroptosis [80]. This is because the resulting c-FLIPL/caspase-8 het- mals but they had major defects in cardiac development and their cells
erodimers, while unable to initiate apoptosis, are able to cleave RIPK1 were resistant to numerous inducers of apoptosis. Therefore, caspase-3
leading to disassembly of the complex. Conversely, c-FLIPs/caspase-8 or caspase-7 can each compensate for the loss of the other during de-
heterodimers prevent the cleavage of RIPK1 by caspase-8 and, there- velopment. These studies also suggest that caspase-3 and -7 rather than
fore, promote the formation of the RIPK1/RIPK3 complex [81]. caspase-6 are the main executioners of apoptosis. Caspase-8 deficient
Although c-FLIPL/S can directly bind to FADD, in vitro DISC recon- mice are also embryonic lethal [104]. However, the phenotype some-
stitution studies have shown that c- FLIPL/S is recruited to the DISC what mirrors the Casp3/Casp7 double knockout, with a failure of the

5
A. Boice and L. Bouchier-Hayes BBA - Molecular Cell Research 1867 (2020) 118688

Fig. 4. The death receptor pathway of apoptosis. At the plasma membrane, the ligands CD95L, TNF-related apoptosis-inducing ligand (TRAIL), and tumor necrosis
factor α (TNFα) induce multimerization and activation of the death receptors CD95, TRAIL R1/R2, and TNFR1 respectively leading to death-inducing signaling
complex (DISC) assembly. The CD95 or TRAILR1/R2 DISC leads to caspase-8 induced proximity via fas-associated protein with death domain (FADD) binding the
intracellular death domain (DD) of the receptor and the death effector domain (DED) of caspase-8. Caspase-8 activates the downstream caspases directly or indirectly
through BID cleavage and mitochondrial outer membrane permeabilization (MOMP). TNFR1 recruits receptor-interacting serine/threonine-protein kinase 1 (RIPK1),
TNFR1-associated death domain protein (TRADD), TNF receptor associated factor 2 (TRAF2), cellular inhibitor of apoptosis protein 1/2 (cIAP1/2), and linear
ubiquitin assembly complex (LUBAC), comprised of HOIL1-interacting protein (HOIP), heme-oxidized IRP2 ubiquitin ligase 1 (HOIL1), and SHANK Associated RH
Domain Interactor (Sharpin). LUBAC adds linear ubiquitin chains (Ub) to RIPK1 and NF-kappa-B essential modulator (NEMO) to recruit the inhibitor of nuclear factor
kappa-B kinase (IKK)complex to activate nuclear factor kappa-light-chain-enhancer of activated B cells (NFκB) (p50/p65). cIAP1/2 induce NIK degradation to inhibit
non-canonical NFκB (RelB/p50). In the absence of cIAP1/2, TRADD/RIPK1/TRAF2 disassociates from the plasma membrane and recruits FADD and caspase-8
(Complex IIa) to induce apoptosis. In the absence of caspase-8, RIPK1 forms a complex with RIPK3 (Complex IIb). RIPK3 induces phosphorylation of mixed lineage
kinase domain like pseudokinase (MLKL), resulting in necroptosis. Cellular-FLICE inhibitory protein long or short (c-FLIPL/s) can form heterodimers with caspase-8.
At high concentrations c-FLIPL and c-FLIPs inhibit caspase-8 activation. c-FLIPL interaction with caspase-8 within complex IIa destabilizes the complex IIb, promoting
survival. For simplicity, c-FLIPs is shown in the DISC and c-FLIPL is shown in complex IIa but the isoforms can interact with both complexes.

cardiovascular system to develop. Deletion of Ripk3 in Casp8 deficient Finally, chemically-induced liver tumors have been reported to grow
mice rescued the embryonic lethality [105]. Therefore, the lethal de- more aggressively in the absence of caspase-2 [112]. Despite this,
velopmental defect of Casp8−/− mice is not a result of disrupting cas- caspase-2 does not appear to act universally as a tumor suppressor. For
pase-8's role in regulating apoptosis but due to its failure to suppress example, deletion of caspase-2 has no effect on the growth or incidence
necroptosis. of chemically-induced fibrosarcoma or irradiation-induced lymphoma
Of the apoptotic caspases, caspase-2 is the only one not to result in a [113]. In a murine model of neuroblastoma, deletion of caspase-2 has
developmental abnormality when deleted in mice. Mice are born actually been reported to delay TH-MYCN-induced neuroblastoma
normal and at the expected Mendelian ratios [106,107]. Although these [114]. This may suggest that caspase-2 has both tumor promoting and
mice do not show any evidence of spontaneous tumorigenesis, caspase- tumor suppressive effects depending on the type of cancer.
2 has been shown to function as a tumor suppressor when the Casp2−/− At present, it is unclear if caspase-2 is unique in its tumor sup-
mice have been crossed to a number of established murine models of pressive abilities or if similar functions of the other caspases are masked
cancer. Compared to wild type mice, caspase-2 deficient mice showed by their essential functions in development. Caspase-1 is another cas-
significantly accelerated tumorigenesis in the Eμ-Myc lymphoma model pase that does not show a lethal phenotype when knocked out [115]
and in ataxia-telangiectasia mutated (Atm)−/− lymphomas [108,109]. and Casp1 deficiency has been shown to enhance tumor formation in an
Caspase-2 has also been shown to have a tumor suppressor role in non- inflammation-induced colorectal cancer model [116]. Conditional
hematologic tumors. Deficiency of caspase-2 significantly accelerated knockout of caspase-8 in B cells resulted in B-cell lymphomagenesis, the
tumorigenesis in the MMTV/c-neu murine mammary carcinogenesis incidence of which was further enhanced by co-deletion of Tp53 [117].
model, and in (LSL)-KrasG12D/+ lung adenocarcinoma [110,111]. Similarly, neural crest specific deletion of Casp8 in TH-MYCN mice led

6
A. Boice and L. Bouchier-Hayes BBA - Molecular Cell Research 1867 (2020) 118688

to more aggressive neuroblastoma [118]. However, in the latter case, rare because of its essential roles in development and other processes.
there was no difference in tumor incidence, latency or size. Rather, loss Instead of direct inactivation of the caspase itself, it is more likely that
of caspase-8 was associated with a significant increased incidence of caspases are compromised in cancer by mutation or deregulation of
bone metastasis and rate of secondary tumors. Based on these ob- upstream regulatory factors impacting caspase activation. Another
servations, it is possible that the use of additional lineage specific de- factor that could explain the contradictory associations of caspase ex-
letions could uncover tumor suppressive roles for the additional cas- pression with cancer incidence or severity in both human populations
pases. and murine models is that many caspases have been shown to have
roles in addition to inducing apoptosis. The availability of distinct
6. What are the consequences of deregulated caspase expression caspase substrate repertoires in different cell types or even disease stage
in cancer? could impact the downstream consequences of caspase activation. This
may explain some of the differences seen in the correlations of caspase
Evidence of significant associations between caspase expression le- expression with prognosis or disease severity. It is thus possible that the
vels and cancer severity has often been either slight or contradictory. specific conditions of the cell in which a caspase is activated determines
For example, in pediatric neuroblastoma, caspase-8 expression was whether that caspase induces apoptosis or an additional non-apoptotic
shown to be absent or deficient in 75% of primary neuroblastoma cases process. This could potentially play a fundamental role in determining
and this downregulation appeared to be due to gene silencing by me- how a caspase can impact cancer. Therefore, it is important to under-
thylation of the CpG rich 5′ flanking region of CASP8 [119,120]. Al- stand both the mechanisms of action of caspases and the downstream
though there was no correlation between caspase-8 protein expression functions of each caspase to be able to fully appreciate their role in
and survival [119], neuroblastoma cell lines with methylated CASP8 cancer and how to effectively target these proteins for therapeutic
were resistant to doxorubicin and CD95L-mediated apoptosis [120]. success.
However, CASP8 methylation has not been reported in any other tumor
type, and in glioblastoma, increased expression of caspase-8 protein 7. Targeting caspase pathways
was reported to be associated with a worse prognosis [121].
Although caspase-2 has been shown to function as a tumor sup- One of the goals of cancer treatment is to design therapies that can
pressor in mouse models, corroborating evidence from human studies remove the cancer cells without harming the healthy tissue.
has been somewhat lacking. CASP2 is located on chromosome 7q34–35 Compounds, such as TRAIL, that have been shown to selectively kill
that is frequently deleted in leukemia. Lower expression of caspase-2 cancer cells [131] could achieve this. Another important challenge is
was associated with drug resistance in childhood leukemia subtypes designing strategies to overcome inherent resistance of tumor cells to
such as acute lymphoblastic leukemia (T-ALL) and acute myeloid leu- conventional therapies. Leveraging the fact that caspase inhibitors in-
kemia (AML) [122], and reduced expression of caspase-2 in AML has cluding XIAP and c-FLIP are often overexpressed in cancer [95,132] can
been correlated with poor prognosis [123]. High levels of inactive provide a means to specifically target tumors.
caspase-2 protein have been linked to a worse prognosis in cancer pa- TRAIL is a death receptor ligand with potential as a cancer therapy
tients with ALL and AML, suggesting that defects in caspase-2 activation because it has been shown in mouse models and in cell lines to selec-
may also contribute to patient outcomes [124,125]. CASP2 is subject to tively kill cancer cells but not primary cells or tissue [131]. In contrast,
epigenetic silencing in a high proportion of mesotheliomas and cases of treatment with CD95L induces a lethal level of hepatocyte cell death in
adult ALL [126]. In contrast, low caspase-2 expression correlates with mice [133]. The receptors for TRAIL, TRAIL-R1 and TRAIL-R2, are
survival in MYCN non-amplified neuroblastoma [114]. This is some- expressed widely in normal tissue [134,135]. TRAIL also has two decoy
what consistent with the data from the mouse model, although, in mice, receptors, TRAIL-R3/DcR1 and TRAIL-R4/DcR2, which do not express
the delay in tumorigenesis associated with caspase-2 loss was observed an intracellular DD and, thus, cannot transduce the apoptotic signal to
in a MYCN amplified context. It must be noted that many of these caspase-8 [136]. An association between TRAIL-R4 expression and
studies were hampered by small sample sizes or the significance of the TRAIL resistance has been reported in breast cancer cell lines [137],
association was not overwhelmingly large. Therefore, the clinical sig- suggesting that the decoy receptors provide a mechanism for the
nificance of caspase-2 expression levels in human cancer has not been cancer-specific killing of TRAIL. However, in other cancers, such as
fully elucidated. multiple myeloma, no correlation between the level of decoy receptor
There are similar conflicting results concerning the validity of cas- expression and the efficiency of TRAIL-mediated killing has been shown
pase-3 as a biomarker for human cancer. In breast tissue samples col- [138]. A number of TRAIL agonists have been developed to clinically
lected from 46 cancer patients, 31 of whom had adenocarcinoma, 75% activate the TRAIL pathway for cancer therapies. These include ago-
of the tumor samples analyzed were found devoid of both caspase-3 nistic antibodies against TRAIL-R1 and soluble recombinant forms of
protein and mRNA, [127]. However, some studies have shown a link TRAIL. Unfortunately, most cancers have demonstrated resistance to
between increased caspase-3 expression and poor cancer prognosis. For these therapies [139]. The resistance factors appear to vary dependent
example, immunohistochemistry for caspase-3 on tissue microarrays on cell type and can be mediated through c-FLIP [140], IAPs [141] or
from a large cohort of early stage breast cancer patients revealed a anti-apoptotic BCL2 family proteins [142]. Combination therapies that
small but statistically significant association between high caspase-3 include TRAIL plus Smac mimetics (see below) or BH3 mimetics (in-
expression and lower rates of cancer-specific survival [128]. Similarly, hibitors of anti-apoptotic BCL2 family proteins) have shown promising
increased levels of cleaved caspase-3, as measured by im- preclinical efficacy (reviewed in [139]).
munohistochemistry analysis, were significantly associated with lower To overcome IAP-associated resistance in cancer, a range of drugs
survival rates in gastric cancer, ovarian cancer, cervical cancer, and called Smac mimetics have been developed to target the IAPs. Smac
colorectal cancer [129]. The association of poor cancer prognosis with mimetics were designed to disrupt the interaction between XIAP and
increased levels of cleaved caspase-3 was also detected in studies on active caspases as a means to induce apoptosis of cancer cells [143,144]
oral tongue squamous cell carcinoma [130]. However, stratification They are chemical derivatives of the peptide sequence AVPI, which is
analysis of the results revealed that, in patients with advanced disease the region of Smac that binds to the BIR2 and BIR3 domains of XIAP
progression (i.e. larger tumor size and metastases to lymph nodes), a [145,146]. Although Smac mimetics were originally designed to target
higher level of cleaved caspase-3 was associated with a better prognosis XIAP, they actually have improved efficiency in targeting cIAP1 and
[130]. cIAP2. When bound to cIAP1/2, Smac mimetics enhance the proteo-
The reasons for these conflicting results are not very clear. As dis- somal degradation of cIAP1/2 by inducing a conformational change
cussed, it may be that complete inactivation of an apoptotic caspase is that results in dimerization of the RING domain and facilitating

7
A. Boice and L. Bouchier-Hayes BBA - Molecular Cell Research 1867 (2020) 118688

autoubiquitination and degradation of the IAPs [147]. This facilitates apoptosis, in part, by downregulating FLIP expression [170]. Similarly,
complex II formation, caspase-8 activation and apoptosis. inhibition of AKT has been shown to overcome resistance to cisplatin-
Smac mimetics can be divided into two categories, monovalent or induced apoptosis that has been associated with c-FLIP overexpression
bivalent compounds. Bivalents are made up of two Smac mimetic motifs [171]. Finally, c-FLIP function can be deregulated by interfering with
(monovalents) that are bridged by a linker. Monovalents have the ad- its ability to bind to caspase-8. It has been demonstrated that mild heat
vantage of being orally bioavailable. Bivalent compounds have higher stress induced aggregation of c-FLIP, leading to the depletion of the
potency and cytotoxicity [143,144,148]. Current Smac mimetics used available cytosolic pool of c-FLIP that could be recruited to the DISC
in precilinical studies or in clinical trials include monovalent com- [172]. Generating chemical inhibitors that would disrupt c-FLIP
pounds, such as LCL161 and Debio1143 [149], and the bivalent com- binding to the DISC or complex II would be ideal. However, a number of
pounds birinapant and AZD5582 [150]. In phase I clinical trials, these challenges to this approach exist. The first is to identify compounds that
Smac mimetics have been shown to be tolerable with minimal severe do not simultaneously inhibit caspase-8. In addition, the mechanisms
side-effects or toxicities [151–153]. However, limited anti-tumor ac- and outcomes of c-FLIP-associated inhibition of cell death are different
tivity was observed in these investigations. Indeed, a phase II clinical for the different isoforms and, as discussed, can vary with c-FLIP con-
trial using birinapant as a treatment for platinum-resistant or refractory centration. Thus c-FLIP inhibitors may have different cellular effects
epithelial ovarian cancer was terminated for lack of clinical benefit depending on the isoform expressed and the relative levels of receptor
despite evidence of consistent cIAP1 protein depletion in tumors [154]. to c-FLIP expression. Further clarification of the mechanism of FLIP-
Although several groups have reported that Smac mimetics can be used mediated caspase-8 inhibition will aid the rational drug design of FLIP-
effectively to kill cancer cells [143,144,148,155], they currently show specific inhibitors.
more clinical promise when used in combination with existing thera-
pies. Smac mimetics used combination with therapeutic agents that 8. Using caspases in gene therapy approaches
target death receptors appear to be the most effective methods for
cancer therapy and their use has enhanced the cytotoxicity of TRAIL The discovery that initiator caspase activation is induced by di-
[156]; CD95L [157] and TNFα [143]. merization has directly led to the development of novel therapeutic
The reason for the limited effectiveness of Smac mimetics as a approaches using an inducible version of caspase-9. This approach was
monotherapy can be partially explained by a feedback loop leading to first used as a safety switch in biological therapies used for the treat-
upregulation of cIAP2. Degradation of cIAP1/2 by Smac-mimetics fa- ment of cancer. When haploidentical stem cell transplant is used for the
cilitates non-canonical NFκB activation, triggering the upregulation of treatment of leukemia, a major clinical side effect is graft-versus-host
TNFα and, consequently, TNFα-mediated cell death [143]. cIAP1 and disease (GvHD), an often life-threatening allergic reaction to the
cIAP2 are functionally redundant, so only one needs to be present for transplanted T cells. To test if removing the cells following transplant
their action [158]. Smac mimetic-induced depletion of cIAP1 can lead could counteract GvHD, donor T cells were engineered to express an
to activation non-canonical NFκB signaling, resulting in increased inducible version of caspase-9 (iCasp9). The catalytic domains of cas-
cIAP2 expression [159]. The de novo expressed cIAP2 is not further pase-9 were fused to the FK-binding domain FKBP, dimerization of
depleted by the Smac mimetic because cIAP2 degradation requires which can be induced by a small, otherwise bioinert, molecule AP1903
cIAP1 [160]. Thus, up-regulation of cIAP2 can lead to resistance to (referred to as chemical inducer of dimerization [CID]). Induced di-
Smac mimetic-induced cell death. Another factor that confers resistance merization of FKBP serves to enforce dimerization and activation of
to Smac mimetics occurs in cancers that results from a chromosomal caspase-9 in the absence of any other upstream signal. To test the ef-
translocation resulting in the cIAP2-MALT1 oncogene. cIAP2-MALT1 is fectiveness of this, T cells from a partially matched donor were en-
the most common translocation in Mucosa-associated lymphoid tissue gineered to express iCasp9 and were adoptively transferred into patients
(MALT) lymphoma and creates a fusion protein of the N-terminal por- with high-risk malignancies who had received a T-cell-depleted he-
tion of cIAP2 lacking the RING domain and the C-terminal caspase-like mopoietic stem cell transplant. Although such T cell depleted trans-
domain of the paracaspase MALT1. This fusion results in stabilized plants produce a low incidence of graft versus host disease, the re-
cIAP2 expression and constitutive canonical NFκB activation con- cipients have a higher risk of infection and relapse of their malignancy
tributing to overall cell survival [161]. Thus, blocking NFκB signaling is since they lack the benefit of T cell control of infectious agents and of
an effective way to overcome resistance to Smac mimetics. For ex- residual host malignant cells. Accordingly, investigators gave iCasp9-
ample, inhibition of IKKβ with the drug BMS-345541 blocked the up- expressing T cells after the transplant with the intent of restoring the
regulation of cIAP2 and sensitized H1299 NSCLC cells to apoptosis beneficial effects of T cells, but this time adding the protection of
induced by the Smac mimetic plus TNFα [159]. iCasp9-mediated T cell killing, should GvHD emerge. Following any
Due to the relationship between c-FLIP overexpression and both evidence of GvHD or transplant rejection, the dimerization agent was
poor patient outcomes and TRAIL resistance, inhibiting or down- administered. This resulted in rapid apoptosis and elimination of more
regulating c-FLIP has long been considered a potential means of over- than 90% of the T cells. Remarkably, GvHD subsided in 24–28 h in the
coming resistance in cancer. While small molecule inhibitors of c-FLIP 4/10 patients who had this complication [173]. The T cell reconstitu-
have been in development, their use as clinical agents has yet to be tion in patients given the dimerization agent after GvHD appeared
published. However, other methods have been used to either decrease unaffected when compared to patients who did not have GvHD and the
c-FLIP levels or inhibit its expression. c-FLIP is a short-lived protein that T cells with iCasp9 in the non-GvHD patients remained for more than
is rapidly turned over by the proteasome [162]. A number of agents 2 years after treatment, showing lack of spontaneous iCasp9 dimeriza-
have been shown to enhance proteasome-mediated degradation of c- tion and the elimination of cells that would result [174]. Further
FLIP to overcome chemoresistance. These include cisplatin, which has murine studies have shown the potential for using this approach to
been shown to induce ubiquitination and degradation of c-FLIP in a remove transplanted induced pluripotent stem cells or mesenchymal
p53-dependent manner. Cisplatin-induces p53 expression and promotes stromal cells (MSC) [175,176].
p53 binding to c-FLIP and ITCH, an E3 ligase and key regulator of FLIP The same inducible caspase-9 approach has been tested as a me-
ubiquitination [163–165]. HDAC inhibitors have similarly been shown chanism for directly targeting solid tumors. In this system, MSC were
to induce FLIP depletion to overcome TRAIL resistance [166]. It has used to deliver an adenoviral-based iCasp9 gene to lung tumors
been proposed that HDAC inhibition disrupts c-FLIP binding to Ku70, [177,178]. Upon infusion into mice or humans, MSC naturally home to
which, in turn, promotes its degradation [167]. c-FLIP transcription is lung vasculature and to lung tumors. This allows the iCasp9 gene to be
regulated by a number of factors including NFκB and AKT [168,169]. directly and specifically delivered to NSCLC. Subsequent treatment
Inhibition of NFκB in combination with TRAIL treatment induces with CID would induce caspase-9 activation and would be expected to

8
A. Boice and L. Bouchier-Hayes BBA - Molecular Cell Research 1867 (2020) 118688

kill the tumor cells. While this approach showed promise in NSCLC cell 9. Non-apoptotic functions of caspases
lines, resistance to CID in some of the lines revealed potential limita-
tions. Two methods have been proposed to overcome this. First, in- An important consideration when targeting caspase pathways for
hibition of the proteasome by treatment with bortezomib was shown to therapeutic benefit in cancer is that many caspases have non-apoptotic
enhance caspase-9-induced apoptosis in resistant lines. This appeared roles and these roles may help or hinder the efficacy of tumor killing
to be due to the inhibition of XIAP-mediated degradation of active depending on the context. For example, caspase-2 has been proposed to
caspases [177]. Second, using conditionally replication competent have a role in cellular proliferation. Caspase-2 deficient cells proliferate
adenovirus to encode iCasp9 provided both an amplification step for at higher rates [108,186] and have evidence of aneuploidy and genomic
inducible caspase-9 expression and an additional mechanism for tumor instability [109,110]. In MMTV/c-neu mice, mammary tumors from
destruction, improving the overall effectiveness of the approach [178]. Casp2−/− mice showed little evidence of altered apoptosis, but dis-
More recent studies have similarly used an adeno-associated virus- played a number of indicators of impaired cell division and genomic
based delivery of iCasp9 to AML in a zebrafish xenograft model and to instability including an increased mitotic index, abnormal mitoses,
hepatocellular carcinoma in a murine xenograft model. When CID was karyomegaly, and multinucleation [110]. Similar evidence of increased
administered systemically a significant reduction in survival and tumor genomic instability was shown in Casp2−/−/Atm−/− lymphomas
size was observed respectively [179,180]. [109]. In the Kras lung tumor model, Casp2−/− lung tumors initially
Caspase-3 has similarly been proposed for use in gene therapy ap- responded well to chemotherapy but the tumors promptly rebounded
proaches. It has been proposed that caspase-3 can be targeted to gastric with increased proliferation compared to the Casp2+/+ mice [111]. In
tumors by selectively delivering the caspase to tumors overexpressing colon cancer, it has been shown that aneuploidy tolerance associated
human epidermal growth factor receptor (HER), an oncogene that is with somatic mutations in Bcl9L is largely a result of the impaired
overexpressed in 10–40% human gastric carcinomas. For this, NIH3T3 ability of BCL9L to induce caspase-2 expression [187]. Caspase-2 has
mouse fibroblasts were engineered to express and secrete a fusion been shown to cleave murine double minute 2 (MDM2) [188], which
protein containing a secretion signal, an anti-HER2 monoclonal anti- has been proposed to be a key effector of the role of caspase-2 in pro-
body fragment that would target HER2 expressing cells, a pseudomonas tecting from polyploidy and aneuploidy [187,189]. MDM2 is an E3 li-
exotoxin A-derived transportation domain fragment, and a con- gase that targets p53 for proteasomal degradation. MDM2 is cleaved by
stitutively active caspase-3 that would induce apoptosis in the targeted caspase-2 at Asp-367, liberating the c-terminal RING domain, thereby
cells (HER-PE-CP3). When cultured in the presence of medium from the preventing ubiquitination and degradation of p53 and promoting p53
engineered cells that contained the secreted active HER-PE-CP3, human stability and function [188]. Thus, in the absence of caspase-2, DNA-
gastric cancer cells underwent apoptosis at the level of approximately damage-induced expression of the p53 target genes Pidd, p21, and
20%. Injection of the recombinant protein producing cells into mice Mdm2 was decreased [110]. This promotion of the p53 pathway may
engrafted with SGC7901 gastric cancer cells resulted in a significant underlie the proposed role of caspase-2 in regulating cell division.
reduction in tumor size and a significant increase in survival. However, caspase-2 has been shown to function in the absence of p53
Importantly, there was no evidence of injury or cell death in the major [190] so any role of caspase-2 on cell cycle in a p53-deficient context
organs, demonstrating the successful and specific delivery of the fusion needs to be investigated.
protein to the tumor and not to the non-cancerous tissues [181]. While the roles of caspase-2 in inducing apoptosis and limiting cell
A major advantage of directly engaging caspase-9 or caspase-3 in proliferation may both be expected to have an overall positive impact
tumor cells is that they initiate apoptosis downstream of MOMP. Thus, on impairing cancer growth, non-apoptotic roles of other caspases have
these engineered cells would theoretically not be affected by over- been shown to have opposing effects on cancer growth and treatment
expression of anti-apoptotic BCL2 family proteins that block MOMP. outcomes. In addition to inducing apoptosis, caspase-3 can impair
These anti-apoptotic proteins include BCL2, BCL-XL, and induced tumor growth and progression by cleaving substrates that do not have a
myeloid leukemia cell differentiation protein (MCL1) and are upregu- direct role in apoptosis. Ras GTPase (RasGAP) is a caspase-3 substrate
lated in many human cancers, providing a key mechanism for resistance that, when cleaved at positions 157 and 455, sensitizes tumor cells to
to chemotherapy (reviewed in [182]). However, it has been shown that apoptosis [191]. However, under conditions of low cellular stress that
caspase-3 activation can induce a positive feedback loop to the mi- lead to a low level of caspase-3 activation, caspase-3 only cleaves
tochondria by cleaving BID [183]. Thus, it is conceivable that these RasGAP at position 455, producing an N-terminal fragment that acti-
caspase-9 and caspase-3-based therapies could induce MOMP to some vates the Ras-PI3K-AKT pathway, promoting survival [191]. The cal-
degree, and this may enhance their cytotoxicity in tumors that do not cium-independent phospholipase A2 (iPLA2) is a caspase-3 substrate
overexpress the anti-apoptotic proteins. Consequently, it is important to that, when cleaved, promotes cellular proliferation [192]. This appears
measure the efficiency of such approaches in the presence and absence to provide a mechanism whereby caspase-3 can trigger tumor cell re-
of high levels of BCL2, BCL-XL and MCL1 to determine their effective- population. Immune compromised mice were engrafted with tumor
ness in a wide array of cancers. The iCasp9 suicide could be further cells in the presence of dying wild type or Casp3−/− mouse embryonic
modified based on our knowledge of the caspase-9 activation pathway. fibroblasts (MEF). The resulting tumors showed a strikingly lower rate
For example, engineering a caspase-9 that is unable to bind to XIAP of growth, caspase-3 activation, and cell proliferation in the presence
could be an effective means to overcome resistance. The use of addi- co-injected cells deficient in caspase-3 [192]. This suggested the cas-
tional initiator caspases in the inducible suicide gene system for cancer pase-3 replete dying cells released a growth signal to boost proliferation
treatment has yet to be investigated. However, a similar approach was of surrounding tumor cells. This signal was identified as cleaved iPLA2.
used for caspase-8 to selectively remove adipocytes and cardiac myo- Inactivation of caspase-3 has been shown to have additional anti-tumor
cytes to develop mouse models of lipoatrophy and heart failure re- effects. In the colon cancer cell line, HCT116, CRISPR/Cas9 knockout of
spectively [184,185]. These studies support the concept that using an caspase-3 resulted in reduced colony forming ability in soft agar and
inducible caspase-8 for the induction of apoptosis could potentially be increased sensitivity to radiation and mitomycin C-induced cell death.
developed for cancer therapeutics. Given the differences in how the Upon engraftment into mice, these cells showed similar rates of tumor
different initiator caspases are activated and their outcomes, it is pos- growth but the caspase-3 deficient tumors responded significantly
sible that exploring the use of caspase-2 or caspase-8 in this system better to radiotherapy with lower rates of lung metastases. The caspase-
could have dramatically different effects on the efficiency of cancer cell 3 deficient cells had reduced expression of N-cadherin, Snail, Slug, and
death in a therapeutic setting. zinc finger E-box-binding homeobox 1 (ZEB1), all proteins known to
promote metastasis [193]. Thus, in certain conditions, caspase-3 ap-
pears to promote metastasis.

9
A. Boice and L. Bouchier-Hayes BBA - Molecular Cell Research 1867 (2020) 118688

As discussed, caspase-8 can exert a pro-survival function by in- mitochondria as a result of mitochondrial inner membrane permeabi-
hibiting RIPK3-dependent necrosis. While this mechanism is usually lization following an apoptotic stimulus has been shown to activate
associated with immune defects, necroptosis of endothelial cells in- cGAS/STING when caspases are blocked or under conditions of min-
duced by tumor cells has been shown to induce metastasis. This process ority MOMP [203,204]. cGAS/STING induces the type I interferon re-
was enhanced when caspase-8 was inactivated in the endothelial cells sponse. By inducing an immune response, this pathway may also have
[194]. It has been shown that under non-apoptotic conditions, caspase- an anti-tumor effect but this has not been directly investigated. To-
8 in transformed cells can potentially contribute to metastasis by in- gether, these studies show that inducing CICD rather than apoptosis
creasing cell motility through the promotion of calpain activation, an may have a therapeutic benefit when treating certain cancers. If CICD
important driver of cytoskeletal remodeling through Rac activation, was more effective at inducing an anti-tumorigenic response, then
lamellipodia formation, and cell adhesion turnover [195]. Thus, dif- blocking caspase activation downstream of MOMP through APAF-1
ferent non-apoptotic functions of caspase-8 may both positively and removal would be expected to result in improved outcomes. Arguing
negatively affect outcomes in cancer. against this is the fact that epigenetic silencing of APAF-1 in melanoma
When the mitochondrial pathway is engaged, MOMP had long been cell lines is associated with increased chemoresistance [205], and loss
considered to be a complete process, with all mitochondria being per- of heterozygosity of APAF-1 did not correlate with changes in overall
meabilized at once [196]. More recent studies have challenged this survival [206]. Nonetheless, it is not clear if this would be the case in all
view, showing that it is possible for only a small subset of the mi- cancer types. While many caspase inhibitors have been patented for use
tochondria in a cell to undergo MOMP. This process has been termed in diseases ranging from stroke to neurodegeneration [207], their ef-
minority MOMP [197]. Minority MOMP has been shown to induce ficacy, if any, in combination with chemotherapy for cancer treatment
small amounts of caspase activation that are not sufficient to induce has not been explored.
apoptosis. Instead, through cleavage of the caspase-3 substrate, ICAD, Glyceraldehyde-3-phosphate dehydrogenase (GAPDH) has been
which is an inhibitor of the DNase CAD, CAD-mediated DNA damage shown to protect from CICD [208]. GAPDH is overexpressed in many
ensues, which can lead to accumulation of genomic instability. The fact human tumors [209] and its overexpression results in accelerated Eμ-
that low levels of caspase activity can result in outcomes that could Myc lymphomagenesis [210]. Overexpression of GAPDH increased
either increase resistance to therapies or enhance tumor growth is an glycolysis and enhanced autophagy by increasing expression of autop-
important consideration when targeting caspase pathways. As dis- hagy related 12 (Atg12). These two processes appear to cooperate to
cussed, in trials using the caspase-9 suicide gene, up to 10% of the cells inhibit CICD [208]. In the absence of caspase activation, GAPDH has
do not undergo apoptosis despite expressing the iCasp9. While it has not also been shown to stabilize the AKT pathway leading to a mildly in-
been investigated whether this small population of cells induced lower creased BCL-XL expression. The increased expression of BCL-XL was
levels of caspase activation or if these cells show any evidence of sufficient to protect a subset of mitochondria from MOMP leading to
genomic instability or pro-inflammatory markers, these events should increased clonogenic outgrowth [211]. This form of incomplete MOMP
be considered when translating these studies to more direct tumor (iMOMP) can occur regardless of the presence or absence of caspase
killing mechanisms. To avoid relapse or developing resistance to these activity, but, when caspases are blocked, there is a strong correlation
therapies, these approaches need to develop to the point that they in- between iMOMP and recovery from CICD and transformation potential
duce maximal apoptosis in the tumors and the biological characteristics [212].
of the cells left behind need to be fully evaluated.

10. Caspase independent cell death 11. Conclusions

Under many circumstances, even if caspases are inhibited, cell death In summary, approaches that directly induce caspase activation or
will still occur by some mechanism. This is especially true of the in- lift the brakes on existing inhibitors of caspase activation represent
trinsic pathway. MOMP is often considered the point of no return in cell promising strategies for the treatment of different types of cancer.
death. For example, in a study of Apaf1-knockout embryonic stem cells, However, certain challenges remain. Since the discovery of the caspase
deficiency in APAF1 was not sufficient to block cell death but inhibition family of proteases, our knowledge of their functions has expanded
of MOMP by overexpression of BCL2 allowed cell survival [198]. Si- from being solely drivers of apoptosis to having multifaceted roles in
milarly, cells lacking both effectors of MOMP, BCL2 associated X cellular biology ranging from regulating cellular proliferation to indu-
apoptosis regulator (BAX) and BCL2-antagonist killer 1 (BAK), are cing genomic instability. Adding to this complexity, different substrate
completely resistant to stimuli that engage the mitochondrial pathway repertoires of individual caspases that change with cell types and dis-
[199]. Thus, if caspases are blocked downstream of MOMP, caspase- ease progression can significantly impact the outcome of activation of a
independent cell death (CICD) will occur, whereas if MOMP is blocked, specific caspase. Because of these diverse roles, insufficient or in-
the cell will survive. CICD is thought to occur primarily as a result of complete activation of caspases may even contribute to chemoresis-
eventual loss of ATP from the mitochondria due to MOMP. However, tance or resurgence of tumor growth. Thus, in order to completely
the release of apoptosis inducing factor (AIF) and endonuclease G leverage the apoptotic caspase pathways for therapeutic or preventative
(EndoG) from mitochondria has also been proposed to play a role. purposes both the mechanisms and full consequences of their activation
EndoG is an apoptotic DNase that, once released from the mitochon- need to be considered and their multiple critical roles in determining
dria, translocates to the nucleus to induce DNA fragmentation in- cell fate need to be fully understood. Great strides have been made in
dependent of caspases [200]. AIF similarly translocates to the nucleus elucidating these biochemical pathways and will ensure that these
to induce DNA fragmentation and peripheral chromatin condensation challenges can be overcome to provide continued improvements in the
in the absence of caspases [201]. However, most documented forms of development and application of new therapeutic strategies to prevent
CICD are not characterized by fragmented nuclei and other mechanisms and cure cancer.
are likely involved.
There are a number of consequences of CICD. Studies have shown
that, in the absence of caspases, both the NFκB and the cGAS/STING Declaration of competing interest
pathways are engaged. These independent pathways result in a proin-
flammatory response that have anti-tumor effects. For example, CICD- The authors declare that they have no known competing financial
associated NFκB activation led to a profound reduction in the growth of interests or personal relationships that could have appeared to influ-
BCL2-dependent tumors [202]. Release of mtDNA from the ence the work reported in this paper.

10
A. Boice and L. Bouchier-Hayes BBA - Molecular Cell Research 1867 (2020) 118688

Acknowledgements [26] H.R. Stennicke, M. Renatus, M. Meldal, G.S. Salvesen, Internally quenched fluor-
escent peptide substrates disclose the subsite preferences of human caspases 1, 3,
6, 7 and 8, Biochem J 350 (Pt 2) (2000) 563–568.
We thank Alexandra Brown, Beatriz Bolivar, and Malcolm Brenner [27] G.P. McStay, G.S. Salvesen, D.R. Green, Overlapping cleavage motif selectivity of
for careful reading of the manuscript and their helpful suggestions. caspases: implications for analysis of apoptotic pathways, Cell Death Differ. 15
Research in the Bouchier-Hayes laboratory is funded by a grant from (2008) 322–331.
[28] B. Cabukusta, N.T. Nettebrock, M. Kol, A. Hilderink, F.G. Tafesse, J.C.M. Holthuis,
the National Institute of General Medical Sciences of the National Ceramide phosphoethanolamine synthase SMSr is a target of caspase-6 during
Institutes of Health under Award Number R01GM121389 and support apoptotic cell death, Biosci. Rep. 37 (2017).
from Texas Children's Cancer and Hematology Centers. [29] S. Shahzidi, A. Brech, M. Sioud, X. Li, Z. Suo, J.M. Nesland, Q. Peng, Lamin A/C
cleavage by caspase-6 activation is crucial for apoptotic induction by photo-
dynamic therapy with hexaminolevulinate in human B-cell lymphoma cells,
Author contributions Cancer Lett. 339 (2013) 25–32.
[30] J.G. Walsh, S.P. Cullen, C. Sheridan, A.U. Luthi, C. Gerner, S.J. Martin,
Executioner caspase-3 and caspase-7 are functionally distinct proteases, Proc. Natl.
AB and LBH drafted and edited the manuscript. AB generated the
Acad. Sci. U. S. A. 105 (2008) 12815–12819.
figures. [31] E.A. Slee, C. Adrain, S.J. Martin, Executioner caspase-3, -6, and -7 perform dis-
tinct, non-redundant roles during the demolition phase of apoptosis, J. Biol. Chem.
References 276 (2001) 7320–7326.
[32] A. Schweizer, C. Briand, M.G. Grutter, Crystal structure of caspase-2, apical in-
itiator of the intrinsic apoptotic pathway, J. Biol. Chem. 278 (2003) 42441–42447.
[1] S.J. Martin, D.R. Green, Protease activation during apoptosis: death by a thousand [33] J. Rotonda, D.W. Nicholson, K.M. Fazil, M. Gallant, Y. Gareau, M. Labelle,
cuts? Cell 82 (1995) 349–352. E.P. Peterson, D.M. Rasper, R. Ruel, J.P. Vaillancourt, N.A. Thornberry,
[2] J.F. Kerr, A.H. Wyllie, A.R. Currie, Apoptosis: a basic biological phenomenon with J.W. Becker, The three-dimensional structure of apopain/CPP32, a key mediator of
wide-ranging implications in tissue kinetics, Br. J. Cancer 26 (1972) 239–257. apoptosis, Nat. Struct. Biol. 3 (1996) 619–625.
[3] S. Arandjelovic, K.S. Ravichandran, Phagocytosis of apoptotic cells in homeostasis, [34] R. Baumgartner, G. Meder, C. Briand, A. Decock, A. D’Arcy, U. Hassiepen,
Nat. Immunol. 16 (2015) 907–917. R. Morse, M. Renatus, The crystal structure of caspase-6, a selective effector of
[4] D.R. McIlwain, T. Berger, T.W. Mak, Caspase functions in cell death and disease, axonal degeneration, Biochem. J. 423 (2009) 429–439.
Cold Spring Harb. Perspect. Biol. 5 (2013) a008656. [35] S.J. Riedl, P. Fuentes-Prior, M. Renatus, N. Kairies, S. Krapp, R. Huber,
[5] C.B. Thompson, Apoptosis in the pathogenesis and treatment of disease, Science G.S. Salvesen, W. Bode, Structural basis for the activation of human procaspase-7,
267 (1995) 1456–1462. Proc. Natl. Acad. Sci. U. S. A. 98 (2001) 14790–14795.
[6] D. Hanahan, R.A. Weinberg, The hallmarks of cancer, Cell 100 (2000) 57–70. [36] H. Blanchard, L. Kodandapani, P.R. Mittl, S.D. Marco, J.F. Krebs, J.C. Wu,
[7] H.M. Ellis, H.R. Horvitz, Genetic control of programmed cell death in the nema- K.J. Tomaselli, M.G. Grutter, The three-dimensional structure of caspase-8: an
tode C. elegans, Cell 44 (1986) 817–829. initiator enzyme in apoptosis, Structure 7 (1999) 1125–1133.
[8] J. Yuan, S. Shaham, S. Ledoux, H.M. Ellis, H.R. Horvitz, The C. elegans cell death [37] E.N. Shiozaki, J. Chai, D.J. Rigotti, S.J. Riedl, P. Li, S.M. Srinivasula, E.S. Alnemri,
gene ced-3 encodes a protein similar to mammalian interleukin-1 beta-converting R. Fairman, Y. Shi, Mechanism of XIAP-mediated inhibition of caspase-9, Mol. Cell
enzyme, Cell 75 (1993) 641–652. 11 (2003) 519–527.
[9] E.S. Alnemri, D.J. Livingston, D.W. Nicholson, G. Salvesen, N.A. Thornberry, [38] R. Ganesan, P.R. Mittl, S. Jelakovic, M.G. Grutter, Extended substrate recognition
W.W. Wong, J. Yuan, Human ICE/CED-3 protease nomenclature, Cell 87 (1996) in caspase-3 revealed by high resolution X-ray structure analysis, J. Mol. Biol. 359
171. (2006) 1378–1388.
[10] K.M. Boatright, M. Renatus, F.L. Scott, S. Sperandio, H. Shin, I.M. Pedersen, [39] Y. Shi, Mechanisms of caspase activation and inhibition during apoptosis, Mol.
J.E. Ricci, W.A. Edris, D.P. Sutherlin, D.R. Green, G.S. Salvesen, A unified model Cell 9 (2002) 459–470.
for apical caspase activation, Mol. Cell 11 (2003) 529–541. [40] N.D. Thomsen, J.T. Koerber, J.A. Wells, Structural snapshots reveal distinct me-
[11] N.P. Walker, R.V. Talanian, K.D. Brady, L.C. Dang, N.J. Bump, C.R. Ferenz, chanisms of procaspase-3 and -7 activation, Proc. Natl. Acad. Sci. U. S. A. 110
S. Franklin, T. Ghayur, M.C. Hackett, L.D. Hammill, et al., Crystal structure of the (2013) 8477–8482.
cysteine protease interleukin-1 beta-converting enzyme: a (p20/p10)2 homo- [41] A.J. Darmon, D.W. Nicholson, R.C. Bleackley, Activation of the apoptotic protease
dimer, Cell 78 (1994) 343–352. CPP32 by cytotoxic T-cell-derived granzyme B, Nature 377 (1995) 446–448.
[12] N.A. Thornberry, T.A. Rano, E.P. Peterson, D.M. Rasper, T. Timkey, M. Garcia- [42] K. Bose, C. Pop, B. Feeney, A.C. Clark, An uncleavable procaspase-3 mutant has a
Calvo, V.M. Houtzager, P.A. Nordstrom, S. Roy, J.P. Vaillancourt, K.T. Chapman, lower catalytic efficiency but an active site similar to that of mature caspase-3,
D.W. Nicholson, A combinatorial approach defines specificities of members of the Biochemistry 42 (2003) 12298–12310.
caspase family and granzyme B. Functional relationships established for key [43] J. Chai, Q. Wu, E. Shiozaki, S.M. Srinivasula, E.S. Alnemri, Y. Shi, Crystal structure
mediators of apoptosis, J. Biol. Chem. 272 (1997) 17907–17911. of a procaspase-7 zymogen: mechanisms of activation and substrate binding, Cell
[13] J.E. Seaman, O. Julien, P.S. Lee, T.J. Rettenmaier, N.D. Thomsen, J.A. Wells, 107 (2001) 399–407.
Cacidases: caspases can cleave after aspartate, glutamate and phosphoserine re- [44] H.R. Stennicke, Q.L. Deveraux, E.W. Humke, J.C. Reed, V.M. Dixit, G.S. Salvesen,
sidues, Cell Death Differ. 23 (2016) 1717–1726. Caspase-9 can be activated without proteolytic processing, J. Biol. Chem. 274
[14] D.W. Nicholson, Caspase structure, proteolytic substrates, and function during (1999) 8359–8362.
apoptotic cell death, Cell Death Differ. 6 (1999) 1028–1042. [46] C. Adrain, E.A. Slee, M.T. Harte, S.J. Martin, Regulation of apoptotic protease
[15] R.K. Graham, D.E. Ehrnhoefer, M.R. Hayden, Caspase-6 and neurodegeneration, activating factor-1 oligomerization and apoptosis by the WD-40 repeat region, J.
Trends Neurosci. 34 (2011) 646–656. Biol. Chem. 274 (1999) 20855–20860.
[16] K. Hofmann, P. Bucher, J. Tschopp, The CARD domain: a new apoptotic signalling [47] Y. Hu, L. Ding, D.M. Spencer, G. Nunez, WD-40 repeat region regulates Apaf-1 self-
motif, Trends Biochem. Sci. 22 (1997) 155–156. association and procaspase-9 activation, J. Biol. Chem. 273 (1998) 33489–33494.
[17] F.C. Kischkel, S. Hellbardt, I. Behrmann, M. Germer, M. Pawlita, P.H. Krammer, [48] S.B. Bratton, G. Walker, S.M. Srinivasula, X.M. Sun, M. Butterworth, E.S. Alnemri,
M.E. Peter, Cytotoxicity-dependent APO-1 (Fas/CD95)-associated proteins form a G.M. Cohen, Recruitment, activation and retention of caspases-9 and -3 by Apaf-1
death-inducing signaling complex (DISC) with the receptor, EMBO J. 14 (1995) apoptosome and associated XIAP complexes, EMBO J. 20 (2001) 998–1009.
5579–5588. [49] S. Malladi, M. Challa-Malladi, H.O. Fearnhead, S.B. Bratton, The Apaf-1*procas-
[18] G. Pan, K. O’Rourke, V.M. Dixit, Caspase-9, Bcl-XL, and Apaf-1 form a ternary pase-9 apoptosome complex functions as a proteolytic-based molecular timer,
complex, J. Biol. Chem. 273 (1998) 5841–5845. EMBO J. 28 (2009) 1916–1925.
[19] A. Tinel, J. Tschopp, The PIDDosome, a protein complex implicated in activation [50] O. Micheau, J. Tschopp, Induction of TNF receptor I-mediated apoptosis via two
of caspase-2 in response to genotoxic stress, Science 304 (2004) 843–846. sequential signaling complexes, Cell 114 (2003) 181–190.
[20] M. Muzio, A.M. Chinnaiyan, F.C. Kischkel, K. O'Rourke, A. Shevchenko, J. Ni, [51] Z. Cai, S. Jitkaew, J. Zhao, H.C. Chiang, S. Choksi, J. Liu, Y. Ward, L.G. Wu,
C. Scaffidi, J.D. Bretz, M. Zhang, R. Gentz, M. Mann, P.H. Krammer, M.E. Peter, Z.G. Liu, Plasma membrane translocation of trimerized MLKL protein is required
V.M. Dixit, FLICE, a novel FADD-homologous ICE/CED-3-like protease, is re- for TNF-induced necroptosis, Nat. Cell Biol. 16 (2014) 55–65.
cruited to the CD95 (Fas/APO-1) death–inducing signaling complex, Cell 85 [52] T.M. Fu, Y. Li, A. Lu, Z. Li, P.R. Vajjhala, A.C. Cruz, D.B. Srivastava, F. DiMaio,
(1996) 817–827. P.A. Penczek, R.M. Siegel, K.J. Stacey, E.H. Egelman, H. Wu, Cryo-EM structure of
[21] P. Li, D. Nijhawan, I. Budihardjo, S.M. Srinivasula, M. Ahmad, E.S. Alnemri, caspase-8 tandem DED filament reveals assembly and regulation mechanisms of
X. Wang, Cytochrome c and dATP-dependent formation of Apaf-1/caspase-9 the death-inducing signaling complex, Mol. Cell 64 (2016) 236–250.
complex initiates an apoptotic protease cascade, Cell 91 (1997) 479–489. [53] K. Schleich, U. Warnken, N. Fricker, S. Ozturk, P. Richter, K. Kammerer,
[22] B.C. Baliga, S.H. Read, S. Kumar, The biochemical mechanism of caspase-2 acti- M. Schnolzer, P.H. Krammer, I.N. Lavrik, Stoichiometry of the CD95 death-indu-
vation, Cell Death Differ. 11 (2004) 1234–1241. cing signaling complex: experimental and modeling evidence for a death effector
[23] K.M. Boatright, C. Deis, J.B. Denault, D.P. Sutherlin, G.S. Salvesen, Activation of domain chain model, Mol. Cell 47 (2012) 306–319.
caspases-8 and -10 by FLIP(L), Biochem. J. 382 (2004) 651–657. [54] L.S. Dickens, R.S. Boyd, R. Jukes-Jones, M.A. Hughes, G.L. Robinson, L. Fairall,
[24] G.S. Salvesen, V.M. Dixit, Caspase activation: the induced-proximity model, Proc. J.W. Schwabe, K. Cain, M. Macfarlane, A death effector domain chain DISC model
Natl. Acad. Sci. U. S. A. 96 (1999) 10964–10967. reveals a crucial role for caspase-8 chain assembly in mediating apoptotic cell
[25] B.E. Bolivar, T.P. Vogel, L. Bouchier-Hayes, Inflammatory caspase regulation: death, Mol. Cell 47 (2012) 291–305.
maintaining balance between inflammation and cell death in health and disease, [55] Y. Guo, S.M. Srinivasula, A. Druilhe, T. Fernandes-Alnemri, E.S. Alnemri, Caspase-
FEBS J. 286 (2019) 2628–2644. 2 induces apoptosis by releasing proapoptotic proteins from mitochondria, J. Biol.

11
A. Boice and L. Bouchier-Hayes BBA - Molecular Cell Research 1867 (2020) 118688

Chem. 277 (2002) 13430–13437. I.N. Lavrik, Long and short isoforms of c-FLIP act as control checkpoints of DED
[56] H. Li, H. Zhu, C.J. Xu, J. Yuan, Cleavage of BID by caspase 8 mediates the mi- filament assembly, Oncogene 39 (2019) 1756–1772.
tochondrial damage in the Fas pathway of apoptosis, Cell 94 (1998) 491–501. [83] K. Schleich, J.H. Buchbinder, S. Pietkiewicz, T. Kahne, U. Warnken, S. Ozturk,
[57] C. Scaffidi, S. Fulda, A. Srinivasan, C. Friesen, F. Li, K.J. Tomaselli, K.M. Debatin, M. Schnolzer, M. Naumann, P.H. Krammer, I.N. Lavrik, Molecular architecture of
P.H. Krammer, M.E. Peter, Two CD95 (APO-1/Fas) signaling pathways, EMBO J. the DED chains at the DISC: regulation of procaspase-8 activation by short DED
17 (1998) 1675–1687. proteins c-FLIP and procaspase-8 prodomain, Cell Death Differ. 23 (2016)
[58] Q.L. Deveraux, R. Takahashi, G.S. Salvesen, J.C. Reed, X-linked IAP is a direct 681–694.
inhibitor of cell-death proteases, Nature 388 (1997) 300–304. [84] D.W. Chang, Z. Xing, Y. Pan, A. Algeciras-Schimnich, B.C. Barnhart, S. Yaish-Ohad,
[59] N. Roy, Q.L. Deveraux, R. Takahashi, G.S. Salvesen, J.C. Reed, The c-IAP-1 and c- M.E. Peter, X. Yang, c-FLIP(L) is a dual function regulator for caspase-8 activation
IAP-2 proteins are direct inhibitors of specific caspases, EMBO J. 16 (1997) and CD95-mediated apoptosis, EMBO J. 21 (2002) 3704–3714.
6914–6925. [85] N. Fricker, J. Beaudouin, P. Richter, R. Eils, P.H. Krammer, I.N. Lavrik, Model-
[60] B.P. Eckelman, G.S. Salvesen, The human anti-apoptotic proteins cIAP1 and cIAP2 based dissection of CD95 signaling dynamics reveals both a pro- and antiapoptotic
bind but do not inhibit caspases, J. Biol. Chem. 281 (2006) 3254–3260. role of c-FLIPL, J. Cell Biol. 190 (2010) 377–389.
[61] S.J. Riedl, M. Renatus, R. Schwarzenbacher, Q. Zhou, C. Sun, S.W. Fesik, [86] T. Fernandes-Alnemri, R.C. Armstrong, J. Krebs, S.M. Srinivasula, L. Wang,
R.C. Liddington, G.S. Salvesen, Structural basis for the inhibition of caspase-3 by F. Bullrich, L.C. Fritz, J.A. Trapani, K.J. Tomaselli, G. Litwack, E.S. Alnemri, In
XIAP, Cell 104 (2001) 791–800. vitro activation of CPP32 and Mch3 by Mch4, a novel human apoptotic cysteine
[62] J. Chai, E. Shiozaki, S.M. Srinivasula, Q. Wu, P. Datta, E.S. Alnemri, Y. Shi, protease containing two FADD-like domains, Proc. Natl. Acad. Sci. U. S. A. 93
Structural basis of caspase-7 inhibition by XIAP, Cell 104 (2001) 769–780. (1996) 7464–7469.
[63] Y. Suzuki, Y. Nakabayashi, R. Takahashi, Ubiquitin-protein ligase activity of X- [87] S. Horn, M.A. Hughes, R. Schilling, C. Sticht, T. Tenev, M. Ploesser, P. Meier,
linked inhibitor of apoptosis protein promotes proteasomal degradation of cas- M.R. Sprick, M. MacFarlane, M. Leverkus, Caspase-10 negatively regulates cas-
pase-3 and enhances its anti-apoptotic effect in Fas-induced cell death, Proc. Natl. pase-8-mediated cell death, switching the response to CD95L in favor of NF-
Acad. Sci. U. S. A. 98 (2001) 8662–8667. kappaB activation and cell survival, Cell Rep. 19 (2017) 785–797.
[64] C. Du, M. Fang, Y. Li, L. Li, X. Wang, Smac, a mitochondrial protein that promotes [88] X. Li, X. Pan, H. Zhang, D. Lei, D. Liu, F. Xu, X. Luan, Overexpression of cFLIP in
cytochrome c-dependent caspase activation by eliminating IAP inhibition, Cell 102 head and neck squamous cell carcinoma and its clinicopathologic correlations, J.
(2000) 33–42. Cancer Res. Clin. Oncol. 134 (2008) 609–615.
[65] A.M. Verhagen, P.G. Ekert, M. Pakusch, J. Silke, L.M. Connolly, G.E. Reid, [89] J.S. Riley, R. Hutchinson, D.G. McArt, N. Crawford, C. Holohan, I. Paul, S. Van
R.L. Moritz, R.J. Simpson, D.L. Vaux, Identification of DIABLO, a mammalian Schaeybroeck, M. Salto-Tellez, P.G. Johnston, D.A. Fennell, K. Gately, K. O’Byrne,
protein that promotes apoptosis by binding to and antagonizing IAP proteins, Cell R. Cummins, E. Kay, P. Hamilton, I. Stasik, D.B. Longley, Prognostic and ther-
102 (2000) 43–53. apeutic relevance of FLIP and procaspase-8 overexpression in non-small cell lung
[66] Y. Suzuki, Y. Imai, H. Nakayama, K. Takahashi, K. Takio, R. Takahashi, A serine cancer, Cell Death Dis. 4 (2013) e951.
protease, HtrA2, is released from the mitochondria and interacts with XIAP, in- [90] G. Valente, F. Manfroi, C. Peracchio, G. Nicotra, R. Castino, G. Nicosia, S. Kerim,
ducing cell death, Mol Cell 8 (2001) 613–621. C. Isidoro, cFLIP expression correlates with tumour progression and patient out-
[67] L.M. Martins, I. Iaccarino, T. Tenev, S. Gschmeissner, N.F. Totty, N.R. Lemoine, come in non-Hodgkin lymphomas of low grade of malignancy, Br. J. Haematol.
J. Savopoulos, C.W. Gray, C.L. Creasy, C. Dingwall, J. Downward, The serine 132 (2006) 560–570.
protease Omi/HtrA2 regulates apoptosis by binding XIAP through a reaper-like [91] M.B. Valnet-Rabier, B. Challier, S. Thiebault, R. Angonin, G. Margueritte,
motif, J. Biol. Chem. 277 (2002) 439–444. C. Mougin, B. Kantelip, E. Deconinck, J.Y. Cahn, T. Fest, c-Flip protein expression
[68] J. Chai, C. Du, J.W. Wu, S. Kyin, X. Wang, Y. Shi, Structural and biochemical basis in Burkitt’s lymphomas is associated with a poor clinical outcome, Br. J. Haematol.
of apoptotic activation by Smac/DIABLO, Nature 406 (2000) 855–862. 128 (2005) 767–773.
[69] S.M. Srinivasula, R. Hegde, A. Saleh, P. Datta, E. Shiozaki, J. Chai, R.A. Lee, [92] D. McLornan, J. Hay, K. McLaughlin, C. Holohan, A.K. Burnett, R.K. Hills,
P.D. Robbins, T. Fernandes-Alnemri, Y. Shi, E.S. Alnemri, A conserved XIAP-in- P.G. Johnston, K.I. Mills, M.F. McMullin, D.B. Longley, A. Gilkes, Prognostic and
teraction motif in caspase-9 and Smac/DIABLO regulates caspase activity and therapeutic relevance of c-FLIP in acute myeloid leukaemia, Br. J. Haematol. 160
apoptosis, Nature 410 (2001) 112–116. (2013) 188–198.
[70] P.J. Jost, S. Grabow, D. Gray, M.D. McKenzie, U. Nachbur, D.C. Huang, P. Bouillet, [93] G.J. Ullenhag, A. Mukherjee, N.F. Watson, A.H. Al-Attar, J.H. Scholefield,
H.E. Thomas, C. Borner, J. Silke, A. Strasser, T. Kaufmann, XIAP discriminates L.G. Durrant, Overexpression of FLIPL is an independent marker of poor prognosis
between type I and type II FAS-induced apoptosis, Nature 460 (2009) 1035–1039. in colorectal cancer patients, Clin. Cancer Res. 13 (2007) 5070–5075.
[71] M. Rothe, M.G. Pan, W.J. Henzel, T.M. Ayres, D.V. Goeddel, The TNFR2-TRAF [94] X.D. Zhou, J.P. Yu, J. Liu, H.S. Luo, H.X. Chen, H.G. Yu, Overexpression of cellular
signaling complex contains two novel proteins related to baculoviral inhibitor of FLICE-inhibitory protein (FLIP) in gastric adenocarcinoma, Clinical science
apoptosis proteins, Cell 83 (1995) 1243–1252. (London, England : 1979) 106 (2004) 397–405.
[72] S. Rahighi, F. Ikeda, M. Kawasaki, M. Akutsu, N. Suzuki, R. Kato, T. Kensche, [95] P. Obexer, M.J. Ausserlechner, X-linked inhibitor of apoptosis protein - a critical
T. Uejima, S. Bloor, D. Komander, F. Randow, S. Wakatsuki, I. Dikic, Specific re- death resistance regulator and therapeutic target for personalized cancer therapy,
cognition of linear ubiquitin chains by NEMO is important for NF-kappaB acti- Front. Oncol. 4 (2014) 197.
vation, Cell 136 (2009) 1098–1109. [96] I. Tamm, S.M. Kornblau, H. Segall, S. Krajewski, K. Welsh, S. Kitada,
[73] T.L. Haas, C.H. Emmerich, B. Gerlach, A.C. Schmukle, S.M. Cordier, E. Rieser, D.A. Scudiero, G. Tudor, Y.H. Qui, A. Monks, M. Andreeff, J.C. Reed, Expression
R. Feltham, J. Vince, U. Warnken, T. Wenger, R. Koschny, D. Komander, J. Silke, and prognostic significance of IAP-family genes in human cancers and myeloid
H. Walczak, Recruitment of the linear ubiquitin chain assembly complex stabilizes leukemias, Clin. Cancer Res. 6 (2000) 1796–1803.
the TNF-R1 signaling complex and is required for TNF-mediated gene induction, [97] C.E. Canman, M.B. Kastan, Induction of apoptosis by tumor suppressor genes and
Mol. Cell 36 (2009) 831–844. oncogenes, Semin. Cancer Biol. 6 (1995) 17–25.
[74] B.J. Zarnegar, Y. Wang, D.J. Mahoney, P.W. Dempsey, H.H. Cheung, J. He, [98] H. Yoshida, Y.Y. Kong, R. Yoshida, A.J. Elia, A. Hakem, R. Hakem, J.M. Penninger,
T. Shiba, X. Yang, W.C. Yeh, T.W. Mak, R.G. Korneluk, G. Cheng, Noncanonical T.W. Mak, Apaf1 is required for mitochondrial pathways of apoptosis and brain
NF-kappaB activation requires coordinated assembly of a regulatory complex of development, Cell 94 (1998) 739–750.
the adaptors cIAP1, cIAP2, TRAF2 and TRAF3 and the kinase NIK, Nat. Immunol. [99] K. Kuida, T.F. Haydar, C.Y. Kuan, Y. Gu, C. Taya, H. Karasuyama, M.S. Su,
9 (2008) 1371–1378. P. Rakic, R.A. Flavell, Reduced apoptosis and cytochrome c-mediated caspase
[75] A. Krueger, I. Schmitz, S. Baumann, P.H. Krammer, S. Kirchhoff, Cellular FLICE- activation in mice lacking caspase 9, Cell 94 (1998) 325–337.
inhibitory protein splice variants inhibit different steps of caspase-8 activation at [100] K. Kuida, T.S. Zheng, S. Na, C. Kuan, D. Yang, H. Karasuyama, P. Rakic,
the CD95 death-inducing signaling complex, J. Biol. Chem. 276 (2001) R.A. Flavell, Decreased apoptosis in the brain and premature lethality in CPP32-
20633–20640. deficient mice, Nature 384 (1996) 368–372.
[76] A. Golks, D. Brenner, C. Fritsch, P.H. Krammer, I.N. Lavrik, C-FLIPR, a new reg- [101] J.R. Leonard, B.J. Klocke, C. D’Sa, R.A. Flavell, K.A. Roth, Strain-dependent neu-
ulator of death receptor-induced apoptosis, J. Biol. Chem. 280 (2005) rodevelopmental abnormalities in caspase-3-deficient mice, J. Neuropathol. Exp.
14507–14513. Neurol. 61 (2002) 673–677.
[77] M.A. Hughes, I.R. Powley, R. Jukes-Jones, S. Horn, M. Feoktistova, L. Fairall, [102] T.S. Zheng, S. Hunot, K. Kuida, T. Momoi, A. Srinivasan, D.W. Nicholson,
J.W. Schwabe, M. Leverkus, K. Cain, M. MacFarlane, Co-operative and hierarchical Y. Lazebnik, R.A. Flavell, Deficiency in caspase-9 or caspase-3 induces compen-
binding of c-FLIP and caspase-8: a unified model defines how c-FLIP isoforms satory caspase activation, Nat. Med. 6 (2000) 1241–1247.
differentially control cell fate, Mol. Cell 61 (2016) 834–849. [103] S.A. Lakhani, A. Masud, K. Kuida, G.A. Porter Jr., C.J. Booth, W.Z. Mehal, I. Inayat,
[78] J.W. Yu, P.D. Jeffrey, Y. Shi, Mechanism of procaspase-8 activation by c-FLIPL, R.A. Flavell, Caspases 3 and 7: key mediators of mitochondrial events of apoptosis,
Proc. Natl. Acad. Sci. U. S. A. 106 (2009) 8169–8174. Science 311 (2006) 847–851.
[79] C. Pop, A. Oberst, M. Drag, B.J. Van Raam, S.J. Riedl, D.R. Green, G.S. Salvesen, [104] E.E. Varfolomeev, M. Schuchmann, V. Luria, N. Chiannilkulchai, J.S. Beckmann,
FLIP(L) induces caspase 8 activity in the absence of interdomain caspase 8 clea- I.L. Mett, D. Rebrikov, V.M. Brodianski, O.C. Kemper, O. Kollet, T. Lapidot,
vage and alters substrate specificity, Biochem. J. 433 (2011) 447–457. D. Soffer, T. Sobe, K.B. Avraham, T. Goncharov, H. Holtmann, P. Lonai,
[80] A. Oberst, C.P. Dillon, R. Weinlich, L.L. McCormick, P. Fitzgerald, C. Pop, D. Wallach, Targeted disruption of the mouse Caspase 8 gene ablates cell death
R. Hakem, G.S. Salvesen, D.R. Green, Catalytic activity of the caspase-8-FLIP(L) induction by the TNF receptors, Fas/Apo1, and DR3 and is lethal prenatally,
complex inhibits RIPK3-dependent necrosis, Nature 471 (2011) 363–367. Immunity 9 (1998) 267–276.
[81] M. Feoktistova, P. Geserick, B. Kellert, D.P. Dimitrova, C. Langlais, M. Hupe, [105] W.J. Kaiser, J.W. Upton, A.B. Long, D. Livingston-Rosanoff, L.P. Daley-Bauer,
K. Cain, M. MacFarlane, G. Hacker, M. Leverkus, cIAPs block Ripoptosome for- R. Hakem, T. Caspary, E.S. Mocarski, RIP3 mediates the embryonic lethality of
mation, a RIP1/caspase-8 containing intracellular cell death complex differentially caspase-8-deficient mice, Nature 471 (2011) 368–372.
regulated by cFLIP isoforms, Mol. Cell 43 (2011) 449–463. [106] L. Bergeron, G.I. Perez, G. Macdonald, L. Shi, Y. Sun, A. Jurisicova, S. Varmuza,
[82] L.K. Hillert, N.V. Ivanisenko, J. Espe, C. Konig, V.A. Ivanisenko, T. Kahne, K.E. Latham, J.A. Flaws, J.C. Salter, H. Hara, M.A. Moskowitz, E. Li, A. Greenberg,

12
A. Boice and L. Bouchier-Hayes BBA - Molecular Cell Research 1867 (2020) 118688

J.L. Tilly, J. Yuan, Defects in regulation of apoptosis in caspase-2-deficient mice, [130] P.F. Liu, Y.C. Hu, B.H. Kang, Y.K. Tseng, P.C. Wu, C.C. Liang, Y.Y. Hou, T.Y. Fu,
Genes Dev. 12 (1998) 1304–1314. H.H. Liou, I.C. Hsieh, L.P. Ger, C.W. Shu, Expression levels of cleaved caspase-3
[107] L.A. O'Reilly, P. Ekert, N. Harvey, V. Marsden, L. Cullen, D.L. Vaux, G. Hacker, and caspase-3 in tumorigenesis and prognosis of oral tongue squamous cell car-
C. Magnusson, M. Pakusch, F. Cecconi, K. Kuida, A. Strasser, D.C. Huang, cinoma, PLoS One 12 (2017) e0180620.
S. Kumar, Caspase-2 is not required for thymocyte or neuronal apoptosis even [131] H. Walczak, R.E. Miller, K. Ariail, B. Gliniak, T.S. Griffith, M. Kubin, W. Chin,
though cleavage of caspase-2 is dependent on both Apaf-1 and caspase-9, Cell J. Jones, A. Woodward, T. Le, C. Smith, P. Smolak, R.G. Goodwin, C.T. Rauch,
Death Differ 9 (2002) 832–841. J.C. Schuh, D.H. Lynch, Tumoricidal activity of tumor necrosis factor-related
[108] L.H. Ho, R. Taylor, L. Dorstyn, D. Cakouros, P. Bouillet, S. Kumar, A tumor sup- apoptosis-inducing ligand in vivo, Nat. Med. 5 (1999) 157–163.
pressor function for caspase-2, Proc. Natl. Acad. Sci. U. S. A. 106 (2009) [132] L. Humphreys, M. Espona-Fiedler, D.B. Longley, FLIP as a therapeutic target in
5336–5341. cancer, FEBS J. 285 (2018) 4104–4123.
[109] J. Puccini, S. Shalini, A.K. Voss, M. Gatei, C.H. Wilson, D.K. Hiwase, M.F. Lavin, [133] J. Ogasawara, R. Watanabe-Fukunaga, M. Adachi, A. Matsuzawa, T. Kasugai,
L. Dorstyn, S. Kumar, Loss of caspase-2 augments lymphomagenesis and enhances Y. Kitamura, N. Itoh, T. Suda, S. Nagata, Lethal effect of the anti-Fas antibody in
genomic instability in Atm-deficient mice, Proc. Natl. Acad. Sci. U. S. A. 110 mice, Nature 364 (1993) 806–809.
(2013) 19920–19925. [134] G. Pan, K. O'Rourke, A.M. Chinnaiyan, R. Gentz, R. Ebner, J. Ni, V.M. Dixit, The
[110] M.J. Parsons, L. McCormick, L. Janke, A. Howard, L. Bouchier-Hayes, D.R. Green, receptor for the cytotoxic ligand TRAIL, Science 276 (1997) 111–113.
Genetic deletion of caspase-2 accelerates MMTV/c-neu-driven mammary carci- [135] H. Walczak, M.A. Degli-Esposti, R.S. Johnson, P.J. Smolak, J.Y. Waugh, N. Boiani,
nogenesis in mice, Cell Death Differ. 20 (2013) 1174–1182. M.S. Timour, M.J. Gerhart, K.A. Schooley, C.A. Smith, R.G. Goodwin, C.T. Rauch,
[111] M.R. Terry, R. Arya, A. Mukhopadhyay, K.C. Berrett, P.M. Clair, B. Witt, TRAIL-R2: a novel apoptosis-mediating receptor for TRAIL, EMBO J. 16 (1997)
M.E. Salama, A. Bhutkar, T.G. Oliver, Caspase-2 impacts lung tumorigenesis and 5386–5397.
chemotherapy response in vivo, Cell Death Differ. 22 (2015) 719–730. [136] J.P. Sheridan, S.A. Marsters, R.M. Pitti, A. Gurney, M. Skubatch, D. Baldwin,
[112] S. Shalini, A. Nikolic, C.H. Wilson, J. Puccini, N. Sladojevic, J. Finnie, L. Dorstyn, L. Ramakrishnan, C.L. Gray, K. Baker, W.I. Wood, A.D. Goddard, P. Godowski,
S. Kumar, Caspase-2 deficiency accelerates chemically induced liver cancer in A. Ashkenazi, Control of TRAIL-induced apoptosis by a family of signaling and
mice, Cell Death Differ. 23 (2016) 1727–1736. decoy receptors, Science 277 (1997) 818–821.
[113] C. Manzl, L. Peintner, G. Krumschnabel, F. Bock, V. Labi, M. Drach, A. Newbold, [137] A.D. Sanlioglu, E. Dirice, C. Aydin, N. Erin, S. Koksoy, S. Sanlioglu, Surface TRAIL
R. Johnstone, A. Villunger, PIDDosome-independent tumor suppression by decoy receptor-4 expression is correlated with TRAIL resistance in MCF7 breast
Caspase-2, Cell Death Differ. 19 (2012) 1722–1732. cancer cells, BMC Cancer 5 (2005) 54.
[114] L. Dorstyn, J. Puccini, A. Nikolic, S. Shalini, C.H. Wilson, M.D. Norris, M. Haber, [138] L.F. Lincz, T.X. Yeh, A. Spencer, TRAIL-induced eradication of primary tumour
S. Kumar, An unexpected role for caspase-2 in neuroblastoma, Cell Death Dis. 5 cells from multiple myeloma patient bone marrows is not related to TRAIL re-
(2014) e1383. ceptor expression or prior chemotherapy, Leukemia 15 (2001) 1650–1657.
[115] K. Kuida, J.A. Lippke, G. Ku, M.W. Harding, D.J. Livingston, M.S. Su, R.A. Flavell, [139] J. Lemke, S. von Karstedt, J. Zinngrebe, H. Walczak, Getting TRAIL back on track
Altered cytokine export and apoptosis in mice deficient in interleukin-1 beta for cancer therapy, Cell Death Differ. 21 (2014) 1350–1364.
converting enzyme, Science 267 (1995) 2000–2003. [140] M. MacFarlane, N. Harper, R.T. Snowden, M.J. Dyer, G.A. Barnett, J.H. Pringle,
[116] B. Hu, E. Elinav, S. Huber, C.J. Booth, T. Strowig, C. Jin, S.C. Eisenbarth, G.M. Cohen, Mechanisms of resistance to TRAIL-induced apoptosis in primary B
R.A. Flavell, Inflammation-induced tumorigenesis in the colon is regulated by cell chronic lymphocytic leukaemia, Oncogene 21 (2002) 6809–6818.
caspase-1 and NLRC4, Proc. Natl. Acad. Sci. U. S. A. 107 (2010) 21635–21640. [141] D. Finlay, M. Vamos, M. Gonzalez-Lopez, R.J. Ardecky, S.R. Ganji, H. Yuan, Y. Su,
[117] A. Hakem, S. El Ghamrasni, G. Maire, B. Lemmers, J. Karaskova, A. Jurisicova, T.R. Cooley, C.T. Hauser, K. Welsh, J.C. Reed, N.D. Cosford, K. Vuori, Small-mo-
O. Sanchez, J. Squire, R. Hakem, Caspase-8 is essential for maintaining chromo- lecule IAP antagonists sensitize cancer cells to TRAIL-induced apoptosis: roles of
somal stability and suppressing B-cell lymphomagenesis, Blood 119 (2012) XIAP and cIAPs, Mol. Cancer Ther. 13 (2014) 5–15.
3495–3502. [142] S. Hinz, A. Trauzold, L. Boenicke, C. Sandberg, S. Beckmann, E. Bayer, H. Walczak,
[118] T. Teitz, M. Inoue, M.B. Valentine, K. Zhu, J.E. Rehg, W. Zhao, D. Finkelstein, H. Kalthoff, H. Ungefroren, Bcl-XL protects pancreatic adenocarcinoma cells
Y.D. Wang, M.D. Johnson, C. Calabrese, M. Rubinstein, R. Hakem, W.A. Weiss, against CD95- and TRAIL-receptor-mediated apoptosis, Oncogene 19 (2000)
J.M. Lahti, Th-MYCN mice with caspase-8 deficiency develop advanced neuro- 5477–5486.
blastoma with bone marrow metastasis, Cancer Res. 73 (2013) 4086–4097. [143] E. Varfolomeev, J.W. Blankenship, S.M. Wayson, A.V. Fedorova, N. Kayagaki,
[119] S. Fulda, C. Poremba, B. Berwanger, S. Hacker, M. Eilers, H. Christiansen, B. Hero, P. Garg, K. Zobel, J.N. Dynek, L.O. Elliott, H.J. Wallweber, J.A. Flygare,
K.M. Debatin, Loss of caspase-8 expression does not correlate with MYCN ampli- W.J. Fairbrother, K. Deshayes, V.M. Dixit, D. Vucic, IAP antagonists induce au-
fication, aggressive disease, or prognosis in neuroblastoma, Cancer Res 66 (2006) toubiquitination of c-IAPs, NF-kappaB activation, and TNFalpha-dependent
10016–10023. apoptosis, Cell 131 (2007) 669–681.
[120] T. Teitz, T. Wei, M.B. Valentine, E.F. Vanin, J. Grenet, V.A. Valentine, F.G. Behm, [144] J.E. Vince, W.W. Wong, N. Khan, R. Feltham, D. Chau, A.U. Ahmed,
A.T. Look, J.M. Lahti, V.J. Kidd, Caspase 8 is deleted or silenced preferentially in C.A. Benetatos, S.K. Chunduru, S.M. Condon, M. McKinlay, R. Brink, M. Leverkus,
childhood neuroblastomas with amplification of MYCN, Nat. Med. 6 (2000) V. Tergaonkar, P. Schneider, B.A. Callus, F. Koentgen, D.L. Vaux, J. Silke, IAP
529–535. antagonists target cIAP1 to induce TNFalpha-dependent apoptosis, Cell 131
[121] G. Fianco, M.P. Mongiardi, A. Levi, T. De Luca, M. Desideri, D. Trisciuoglio, D. Del (2007) 682–693.
Bufalo, I. Cina, A. Di Benedetto, M. Mottolese, A. Gentile, D. Centonze, F. Ferre, [145] Y. Huang, R.L. Rich, D.G. Myszka, H. Wu, Requirement of both the second and
D. Barila, Caspase-8 contributes to angiogenesis and chemotherapy resistance in third BIR domains for the relief of X-linked inhibitor of apoptosis protein (XIAP)-
glioblastoma, eLife 6 (2017). mediated caspase inhibition by Smac, J. Biol. Chem. 278 (2003) 49517–49522.
[122] A. Holleman, M.L. den Boer, K.M. Kazemier, H.B. Beverloo, A.R. von Bergh, [146] Z. Liu, C. Sun, E.T. Olejniczak, R.P. Meadows, S.F. Betz, T. Oost, J. Herrmann,
G.E. Janka-Schaub, R. Pieters, Decreased PARP and procaspase-2 protein levels are J.C. Wu, S.W. Fesik, Structural basis for binding of Smac/DIABLO to the XIAP BIR3
associated with cellular drug resistance in childhood acute lymphoblastic leu- domain, Nature 408 (2000) 1004–1008.
kemia, Blood 106 (2005) 1817–1823. [147] E.C. Dueber, A.J. Schoeffler, A. Lingel, J.M. Elliott, A.V. Fedorova, A.M. Giannetti,
[123] S. Dawar, N.H. Shahrin, N. Sladojevic, R.J. D’Andrea, L. Dorstyn, D.K. Hiwase, K. Zobel, B. Maurer, E. Varfolomeev, P. Wu, H.J. Wallweber, S.G. Hymowitz,
S. Kumar, Impaired haematopoietic stem cell differentiation and enhanced K. Deshayes, D. Vucic, W.J. Fairbrother, Antagonists induce a conformational
skewing towards myeloid progenitors in aged caspase-2-deficient mice, Cell Death change in cIAP1 that promotes autoubiquitination, Science 334 (2011) 376–380.
Dis. 7 (2016) e2509. [148] B.L. Probst, L. Liu, V. Ramesh, L. Li, H. Sun, J.D. Minna, L. Wang, Smac mimetics
[124] Z. Estrov, P.F. Thall, M. Talpaz, E.H. Estey, H.M. Kantarjian, M. Andreeff, increase cancer cell response to chemotherapeutics in a TNF-alpha-dependent
D. Harris, Q. Van, M. Walterscheid, S.M. Kornblau, Caspase 2 and caspase 3 manner, Cell Death Differ. 17 (2010) 1645–1654.
protein levels as predictors of survival in acute myelogenous leukemia, Blood 92 [149] Q. Cai, H. Sun, Y. Peng, J. Lu, Z. Nikolovska-Coleska, D. McEachern, L. Liu, S. Qiu,
(1998) 3090–3097. C.Y. Yang, R. Miller, H. Yi, T. Zhang, D. Sun, S. Kang, M. Guo, L. Leopold, D. Yang,
[125] S. Faderl, P.F. Thall, H.M. Kantarjian, M. Talpaz, D. Harris, Q. Van, M. Beran, S. Wang, A potent and orally active antagonist (SM-406/AT-406) of multiple in-
S.M. Kornblau, S. Pierce, Z. Estrov, Caspase 2 and caspase 3 as predictors of hibitor of apoptosis proteins (IAPs) in clinical development for cancer treatment,
complete remission and survival in adults with acute lymphoblastic leukemia, J. Med. Chem. 54 (2011) 2714–2726.
Clin. Cancer Res. 5 (1999) 4041–4047. [150] E.J. Hennessy, A. Adam, B.M. Aquila, L.M. Castriotta, D. Cook, M. Hattersley,
[126] B.C. Christensen, C.J. Marsit, E.A. Houseman, J.J. Godleski, J.L. Longacker, A.W. Hird, C. Huntington, V.M. Kamhi, N.M. Laing, D. Li, T. MacIntyre, C.A. Omer,
S. Zheng, R.F. Yeh, M.R. Wrensch, J.L. Wiemels, M.R. Karagas, R. Bueno, V. Oza, T. Patterson, G. Repik, M.T. Rooney, J.C. Saeh, L. Sha, M.M. Vasbinder,
D.J. Sugarbaker, H.H. Nelson, J.K. Wiencke, K.T. Kelsey, Differentiation of lung H. Wang, D. Whitston, Discovery of a novel class of dimeric Smac mimetics as
adenocarcinoma, pleural mesothelioma, and nonmalignant pulmonary tissues potent IAP antagonists resulting in a clinical candidate for the treatment of cancer
using DNA methylation profiles, Cancer Res. 69 (2009) 6315–6321. (AZD5582), J. Med. Chem. 56 (2013) 9897–9919.
[127] E. Devarajan, A.A. Sahin, J.S. Chen, R.R. Krishnamurthy, N. Aggarwal, A.M. Brun, [151] R.K. Amaravadi, R.J. Schilder, L.P. Martin, M. Levin, M.A. Graham, D.E. Weng,
A. Sapino, F. Zhang, D. Sharma, X.H. Yang, A.D. Tora, K. Mehta, Down-regulation A.A. Adjei, A phase I study of the SMAC-mimetic birinapant in adults with re-
of caspase 3 in breast cancer: a possible mechanism for chemoresistance, fractory solid tumors or lymphoma, Mol. Cancer Ther. 14 (2015) 2569–2575.
Oncogene 21 (2002) 8843–8851. [152] J.R. Infante, E.C. Dees, A.J. Olszanski, S.V. Dhuria, S. Sen, S. Cameron, R.B. Cohen,
[128] X. Pu, S.J. Storr, Y. Zhang, E.A. Rakha, A.R. Green, I.O. Ellis, S.G. Martin, Caspase- Phase I dose-escalation study of LCL161, an oral inhibitor of apoptosis proteins
3 and caspase-8 expression in breast cancer: caspase-3 is associated with survival, inhibitor, in patients with advanced solid tumors, Journal of clinical oncology :
Apoptosis 22 (2017) 357–368. official journal of the American Society of Clinical Oncology 32 (2014)
[129] Q. Hu, J. Peng, W. Liu, X. He, L. Cui, X. Chen, M. Yang, H. Liu, S. Liu, H. Wang, 3103–3110.
Elevated cleaved caspase-3 is associated with shortened overall survival in several [153] H.I. Hurwitz, D.C. Smith, H.C. Pitot, J.M. Brill, R. Chugh, E. Rouits, J. Rubin,
cancer types, Int. J. Clin. Exp. Pathol. 7 (2014) 5057–5070. J. Strickler, G. Vuagniaux, J.M. Sorensen, C. Zanna, Safety, pharmacokinetics, and

13
A. Boice and L. Bouchier-Hayes BBA - Molecular Cell Research 1867 (2020) 118688

pharmacodynamic properties of oral DEBIO1143 (AT-406) in patients with ad- G. Dotti, An inducible caspase 9 suicide gene to improve the safety of mesench-
vanced cancer: results of a first-in-man study, Cancer Chemother. Pharmacol. 75 ymal stromal cell therapies, Stem Cells 28 (2010) 1107–1115.
(2015) 851–859. [177] M. Ando, V. Hoyos, S. Yagyu, W. Tao, C.A. Ramos, G. Dotti, M.K. Brenner,
[154] A.M. Noonan, K.P. Bunch, J.Q. Chen, M.A. Herrmann, J.M. Lee, E.C. Kohn, L. Bouchier-Hayes, Bortezomib sensitizes non-small cell lung cancer to mesench-
C.C. O'Sullivan, E. Jordan, N. Houston, N. Takebe, R.J. Kinders, L. Cao, C.J. Peer, ymal stromal cell-delivered inducible caspase-9-mediated cytotoxicity, Cancer
W.D. Figg, C.M. Annunziata, Pharmacodynamic markers and clinical results from Gene Ther. 21 (2014) 472–482.
the phase 2 study of the SMAC mimetic birinapant in women with relapsed pla- [178] V. Hoyos, F. Del Bufalo, S. Yagyu, M. Ando, G. Dotti, M. Suzuki, L. Bouchier-Hayes,
tinum-resistant or -refractory epithelial ovarian cancer, Cancer 122 (2016) R. Alemany, M.K. Brenner, Mesenchymal stromal cells for linked delivery of on-
588–597. colytic and apoptotic adenoviruses to non-small-cell lung cancers, Mol. Ther. 23
[155] S.L. Petersen, L. Wang, A. Yalcin-Chin, L. Li, M. Peyton, J. Minna, P. Harran, (2015) 1497–1506.
X. Wang, Autocrine TNFalpha signaling renders human cancer cells susceptible to [179] N. Khan, S. Bammidi, G.R. Jayandharan, A CD33 antigen-targeted AAV6 vector
Smac-mimetic-induced apoptosis, Cancer Cell 12 (2007) 445–456. expressing an inducible caspase-9 suicide gene is therapeutic in a xeno-
[156] M. Fakler, S. Loeder, M. Vogler, K. Schneider, I. Jeremias, K.M. Debatin, S. Fulda, transplantation model of acute myeloid leukemia, Bioconjug. Chem. 30 (2019)
Small molecule XIAP inhibitors cooperate with TRAIL to induce apoptosis in 2404–2416.
childhood acute leukemia cells and overcome Bcl-2-mediated resistance, Blood [180] N. Khan, S. Bammidi, S. Chattopadhyay, G.R. Jayandharan, Combination suicide
113 (2009) 1710–1722. gene delivery with an adeno-associated virus vector encoding inducible caspase-9
[157] P. Geserick, M. Hupe, M. Moulin, W.W. Wong, M. Feoktistova, B. Kellert, and a chemical inducer of dimerization is effective in a xenotransplantation model
H. Gollnick, J. Silke, M. Leverkus, Cellular IAPs inhibit a cryptic CD95-induced cell of hepatocellular carcinoma, Bioconjug. Chem. 30 (2019) 1754–1762.
death by limiting RIP1 kinase recruitment, J. Cell Biol. 187 (2009) 1037–1054. [181] D.X. Zhang, P.T. Zhao, L. Xia, L.L. Liu, J. Liang, H.H. Zhai, H.B. Zhang, X.G. Guo,
[158] D.J. Mahoney, H.H. Cheung, R.L. Mrad, S. Plenchette, C. Simard, E. Enwere, K.C. Wu, Y.M. Xu, L.T. Jia, A.G. Yang, S.Y. Chen, D.M. Fan, Potent inhibition of
V. Arora, T.W. Mak, E.C. Lacasse, J. Waring, R.G. Korneluk, Both cIAP1 and cIAP2 human gastric cancer by HER2-directed induction of apoptosis with anti-HER2
regulate TNFalpha-mediated NF-kappaB activation, Proc. Natl. Acad. Sci. U. S. A. antibody and caspase-3 fusion protein, Gut 59 (2010) 292–299.
105 (2008) 11778–11783. [182] A.R. Delbridge, S. Grabow, A. Strasser, D.L. Vaux, Thirty years of BCL-2: trans-
[159] S.L. Petersen, M. Peyton, J.D. Minna, X. Wang, Overcoming cancer cell resistance lating cell death discoveries into novel cancer therapies, Nat. Rev. Cancer 16
to Smac mimetic induced apoptosis by modulating cIAP-2 expression, Proc. Natl. (2016) 99–109.
Acad. Sci. U. S. A. 107 (2010) 11936–11941. [183] E.A. Slee, S.A. Keogh, S.J. Martin, Cleavage of BID during cytotoxic drug and UV
[160] M. Darding, R. Feltham, T. Tenev, K. Bianchi, C. Benetatos, J. Silke, P. Meier, radiation-induced apoptosis occurs downstream of the point of Bcl-2 action and is
Molecular determinants of Smac mimetic induced degradation of cIAP1 and catalysed by caspase-3: a potential feedback loop for amplification of apoptosis-
cIAP2, Cell Death Differ. 18 (2011) 1376–1386. associated mitochondrial cytochrome c release, Cell Death Differ. 7 (2000)
[161] E. Varfolomeev, S.M. Wayson, V.M. Dixit, W.J. Fairbrother, D. Vucic, The inhibitor 556–565.
of apoptosis protein fusion c-IAP2.MALT1 stimulates NF-kappaB activation in- [184] U.B. Pajvani, M.E. Trujillo, T.P. Combs, P. Iyengar, L. Jelicks, K.A. Roth,
dependently of TRAF1 AND TRAF2, J Biol Chem 281 (2006) 29022–29029. R.N. Kitsis, P.E. Scherer, Fat apoptosis through targeted activation of caspase 8: a
[162] M. Poukkula, A. Kaunisto, V. Hietakangas, K. Denessiouk, T. Katajamaki, new mouse model of inducible and reversible lipoatrophy, Nat. Med. 11 (2005)
M.S. Johnson, L. Sistonen, J.E. Eriksson, Rapid turnover of c-FLIPshort is de- 797–803.
termined by its unique C-terminal tail, J. Biol. Chem. 280 (2005) 27345–27355. [185] D. Wencker, M. Chandra, K. Nguyen, W. Miao, S. Garantziotis, S.M. Factor,
[163] M.R. Abedini, E.J. Muller, J. Brun, R. Bergeron, D.A. Gray, B.K. Tsang, Cisplatin J. Shirani, R.C. Armstrong, R.N. Kitsis, A mechanistic role for cardiac myocyte
induces p53-dependent FLICE-like inhibitory protein ubiquitination in ovarian apoptosis in heart failure, J. Clin. Invest. 111 (2003) 1497–1504.
cancer cells, Cancer Res. 68 (2008) 4511–4517. [186] K. Ren, J. Lu, A. Porollo, C. Du, Tumor-suppressing function of caspase-2 requires
[164] J.H. Song, D.K. Song, M. Herlyn, K.C. Petruk, C. Hao, Cisplatin down-regulation of catalytic site Cys-320 and site Ser-139 in mice, J. Biol. Chem. 287 (2012)
cellular Fas-associated death domain-like interleukin-1beta-converting enzyme- 14792–14802.
like inhibitory proteins to restore tumor necrosis factor-related apoptosis-inducing [187] C. Lopez-Garcia, L. Sansregret, E. Domingo, N. McGranahan, S. Hobor,
ligand-induced apoptosis in human melanoma cells, Clin. Cancer Res. 9 (2003) N.J. Birkbak, S. Horswell, E. Gronroos, F. Favero, A.J. Rowan, N. Matthews,
4255–4266. S. Begum, B. Phillimore, R. Burrell, D. Oukrif, B. Spencer-Dene, M. Kovac,
[165] L. Chang, H. Kamata, G. Solinas, J.L. Luo, S. Maeda, K. Venuprasad, Y.C. Liu, G. Stamp, A. Stewart, H. Danielsen, M. Novelli, I. Tomlinson, C. Swanton, BCL9L
M. Karin, The E3 ubiquitin ligase itch couples JNK activation to TNFalpha-induced dysfunction impairs caspase-2 expression permitting aneuploidy tolerance in col-
cell death by inducing c-FLIP(L) turnover, Cell 124 (2006) 601–613. orectal cancer, Cancer Cell 31 (2017) 79–93.
[166] I. Venza, M. Visalli, R. Oteri, D. Teti, M. Venza, Class I-specific histone deacetylase [188] T.G. Oliver, E. Meylan, G.P. Chang, W. Xue, J.R. Burke, T.J. Humpton, D. Hubbard,
inhibitor MS-275 overrides TRAIL-resistance in melanoma cells by downregulating A. Bhutkar, T. Jacks, Caspase-2-mediated cleavage of Mdm2 creates a p53-induced
c-FLIP, Int. Immunopharmacol. 21 (2014) 439–446. positive feedback loop, Mol. Cell 43 (2011) 57–71.
[167] E. Kerr, C. Holohan, K.M. McLaughlin, J. Majkut, S. Dolan, K. Redmond, J. Riley, [189] L.L. Fava, F. Schuler, V. Sladky, M.D. Haschka, C. Soratroi, L. Eiterer, E. Demetz,
K. McLaughlin, I. Stasik, M. Crudden, S. Van Schaeybroeck, C. Fenning, G. Weiss, S. Geley, E.A. Nigg, A. Villunger, The PIDDosome activates p53 in re-
R. O’Connor, P. Kiely, M. Sgobba, D. Haigh, P.G. Johnston, D.B. Longley, sponse to supernumerary centrosomes, Genes Dev. 31 (2017) 34–45.
Identification of an acetylation-dependant Ku70/FLIP complex that regulates FLIP [190] S. Sidi, T. Sanda, R.D. Kennedy, A.T. Hagen, C.A. Jette, R. Hoffmans, J. Pascual,
expression and HDAC inhibitor-induced apoptosis, Cell Death Differ. 19 (2012) S. Imamura, S. Kishi, J.F. Amatruda, J.P. Kanki, D.R. Green, A.A. D'Andrea,
1317–1327. A.T. Look, Chk1 suppresses a caspase-2 apoptotic response to DNA damage that
[168] D.J. Panka, T. Mano, T. Suhara, K. Walsh, J.W. Mier, Phosphatidylinositol 3-ki- bypasses p53, Bcl-2, and caspase-3, Cell 133 (2008) 864–877.
nase/Akt activity regulates c-FLIP expression in tumor cells, J. Biol. Chem. 276 [191] J.Y. Yang, J. Walicki, D. Michod, G. Dubuis, C. Widmann, Impaired Akt activity
(2001) 6893–6896. down-modulation, caspase-3 activation, and apoptosis in cells expressing a cas-
[169] O. Micheau, S. Lens, O. Gaide, K. Alevizopoulos, J. Tschopp, NF-kappaB signals pase-resistant mutant of RasGAP at position 157, Mol. Biol. Cell 16 (2005)
induce the expression of c-FLIP, Mol. Cell. Biol. 21 (2001) 5299–5305. 3511–3520.
[170] T.S. Jani, J. DeVecchio, T. Mazumdar, A. Agyeman, J.A. Houghton, Inhibition of [192] Q. Huang, F. Li, X. Liu, W. Li, W. Shi, F.F. Liu, B. O’Sullivan, Z. He, Y. Peng,
NF-kappaB signaling by quinacrine is cytotoxic to human colon carcinoma cell A.C. Tan, L. Zhou, J. Shen, G. Han, X.J. Wang, J. Thorburn, A. Thorburn,
lines and is synergistic in combination with tumor necrosis factor-related apop- A. Jimeno, D. Raben, J.S. Bedford, C.Y. Li, Caspase 3-mediated stimulation of
tosis-inducing ligand (TRAIL) or oxaliplatin, J. Biol. Chem. 285 (2010) tumor cell repopulation during cancer radiotherapy, Nat. Med. 17 (2011)
19162–19172. 860–866.
[171] M.R. Abedini, E.J. Muller, R. Bergeron, D.A. Gray, B.K. Tsang, Akt promotes [193] M. Zhou, X. Liu, Z. Li, Q. Huang, F. Li, C.Y. Li, Caspase-3 regulates the migration,
chemoresistance in human ovarian cancer cells by modulating cisplatin-induced, invasion and metastasis of colon cancer cells, Int. J. Cancer 143 (2018) 921–930.
p53-dependent ubiquitination of FLICE-like inhibitory protein, Oncogene 29 [194] B. Strilic, L. Yang, J. Albarran-Juarez, L. Wachsmuth, K. Han, U.C. Muller,
(2010) 11–25. M. Pasparakis, S. Offermanns, Tumour-cell-induced endothelial cell necroptosis
[172] A. Morle, C. Garrido, O. Micheau, Hyperthermia restores apoptosis induced by via death receptor 6 promotes metastasis, Nature 536 (2016) 215–218.
death receptors through aggregation-induced c-FLIP cytosolic depletion, Cell [195] B. Helfer, B.C. Boswell, D. Finlay, A. Cipres, K. Vuori, T. Bong Kang, D. Wallach,
Death Dis. 6 (2015) e1633. A. Dorfleutner, J.M. Lahti, D.C. Flynn, S.M. Frisch, Caspase-8 promotes cell mo-
[173] A. Di Stasi, S.K. Tey, G. Dotti, Y. Fujita, A. Kennedy-Nasser, C. Martinez, tility and calpain activity under nonapoptotic conditions, Cancer Res. 66 (2006)
K. Straathof, E. Liu, A.G. Durett, B. Grilley, H. Liu, C.R. Cruz, B. Savoldo, A.P. Gee, 4273–4278.
J. Schindler, R.A. Krance, H.E. Heslop, D.M. Spencer, C.M. Rooney, M.K. Brenner, [196] J.C. Goldstein, N.J. Waterhouse, P. Juin, G.I. Evan, D.R. Green, The coordinate
Inducible apoptosis as a safety switch for adoptive cell therapy, N. Engl. J. Med. release of cytochrome c during apoptosis is rapid, complete and kinetically in-
365 (2011) 1673–1683. variant, Nat. Cell Biol. 2 (2000) 156–162.
[174] X. Zhou, A. Di Stasi, S.K. Tey, R.A. Krance, C. Martinez, K.S. Leung, A.G. Durett, [197] G. Ichim, J. Lopez, S.U. Ahmed, N. Muthalagu, E. Giampazolias, M.E. Delgado,
M.F. Wu, H. Liu, A.M. Leen, B. Savoldo, Y.F. Lin, B.J. Grilley, A.P. Gee, M. Haller, J.S. Riley, S.M. Mason, D. Athineos, M.J. Parsons, B. van de Kooij,
D.M. Spencer, C.M. Rooney, H.E. Heslop, M.K. Brenner, G. Dotti, Long-term out- L. Bouchier-Hayes, A.J. Chalmers, R.W. Rooswinkel, A. Oberst, K. Blyth, M. Rehm,
come after haploidentical stem cell transplant and infusion of T cells expressing D.J. Murphy, S.W.G. Tait, Limited mitochondrial permeabilization causes DNA
the inducible caspase 9 safety transgene, Blood 123 (2014) 3895–3905. damage and genomic instability in the absence of cell death, Mol. Cell 57 (2015)
[175] S. Yagyu, V. Hoyos, F. Del Bufalo, M.K. Brenner, An inducible caspase-9 suicide 860–872.
gene to improve the safety of therapy using human induced pluripotent stem cells, [198] M. Haraguchi, S. Torii, S. Matsuzawa, Z. Xie, S. Kitada, S. Krajewski, H. Yoshida,
Mol. Ther. 23 (2015) 1475–1485. T.W. Mak, J.C. Reed, Apoptotic protease activating factor 1 (Apaf-1)-independent
[176] C.A. Ramos, Z. Asgari, E. Liu, E. Yvon, H.E. Heslop, C.M. Rooney, M.K. Brenner, cell death suppression by Bcl-2, J Exp Med 191 (2000) 1709–1720.

14
A. Boice and L. Bouchier-Hayes BBA - Molecular Cell Research 1867 (2020) 118688

[199] T. Lindsten, A.J. Ross, A. King, W.X. Zong, J.C. Rathmell, H.A. Shiels, E. Ulrich, S.W. Lowe, Inactivation of the apoptosis effector Apaf-1 in malignant melanoma,
K.G. Waymire, P. Mahar, K. Frauwirth, Y. Chen, M. Wei, V.M. Eng, D.M. Adelman, Nature 409 (2001) 207–211.
M.C. Simon, A. Ma, J.A. Golden, G. Evan, S.J. Korsmeyer, G.R. MacGregor, [206] A. Fujimoto, H. Takeuchi, B. Taback, E.C. Hsueh, D. Elashoff, D.L. Morton,
C.B. Thompson, The combined functions of proapoptotic Bcl-2 family members D.S. Hoon, Allelic imbalance of 12q22-23 associated with APAF-1 locus correlates
bak and bax are essential for normal development of multiple tissues, Mol. Cell 6 with poor disease outcome in cutaneous melanoma, Cancer Res. 64 (2004)
(2000) 1389–1399. 2245–2250.
[200] L.Y. Li, X. Luo, X. Wang, Endonuclease G is an apoptotic DNase when released [207] H. Lee, E.A. Shin, J.H. Lee, D. Ahn, C.G. Kim, J.H. Kim, S.H. Kim, Caspase in-
from mitochondria, Nature 412 (2001) 95–99. hibitors: a review of recently patented compounds (2013–2015), Expert Opinion
[201] S.A. Susin, E. Daugas, L. Ravagnan, K. Samejima, N. Zamzami, M. Loeffler, on Therapeutic Patents 28 (2018) 47–59.
P. Costantini, K.F. Ferri, T. Irinopoulou, M.C. Prevost, G. Brothers, T.W. Mak, [208] A. Colell, J.E. Ricci, S. Tait, S. Milasta, U. Maurer, L. Bouchier-Hayes, P. Fitzgerald,
J. Penninger, W.C. Earnshaw, G. Kroemer, Two distinct pathways leading to nu- A. Guio-Carrion, N.J. Waterhouse, C.W. Li, B. Mari, P. Barbry, D.D. Newmeyer,
clear apoptosis, J. Exp. Med. 192 (2000) 571–580. H.M. Beere, D.R. Green, GAPDH and autophagy preserve survival after apoptotic
[202] E. Giampazolias, B. Zunino, S. Dhayade, F. Bock, C. Cloix, K. Cao, A. Roca, cytochrome c release in the absence of caspase activation, Cell 129 (2007)
J. Lopez, G. Ichim, E. Proics, C. Rubio-Patino, L. Fort, N. Yatim, E. Woodham, 983–997.
S. Orozco, L. Taraborrelli, N. Peltzer, D. Lecis, L. Machesky, H. Walczak, [209] B. Altenberg, K.O. Greulich, Genes of glycolysis are ubiquitously overexpressed in
M.L. Albert, S. Milling, A. Oberst, J.E. Ricci, K.M. Ryan, K. Blyth, S.W.G. Tait, 24 cancer classes, Genomics 84 (2004) 1014–1020.
Mitochondrial permeabilization engages NF-kappaB-dependent anti-tumour ac- [210] J. Chiche, S. Pommier, M. Beneteau, L. Mondragon, O. Meynet, B. Zunino,
tivity under caspase deficiency, Nat. Cell Biol. 19 (2017) 1116–1129. A. Mouchotte, E. Verhoeyen, M. Guyot, G. Pages, N. Mounier, V. Imbert,
[203] J.S. Riley, G. Quarato, C. Cloix, J. Lopez, J. O'Prey, M. Pearson, J. Chapman, P. Colosetti, D. Goncalves, S. Marchetti, J. Briere, M. Carles, C. Thieblemont,
H. Sesaki, L.M. Carlin, J.F. Passos, A.P. Wheeler, A. Oberst, K.M. Ryan, S.W. Tait, J.E. Ricci, GAPDH enhances the aggressiveness and the vascularization of non-
Mitochondrial inner membrane permeabilisation enables mtDNA release during Hodgkin's B lymphomas via NF-kappaB-dependent induction of HIF-1alpha,
apoptosis, EMBO J 37 (2018). Leukemia 29 (2015) 1163–1176.
[204] D. Brokatzky, B. Dorflinger, A. Haimovici, A. Weber, S. Kirschnek, J. Vier, A. Metz, [211] M.A. Jacquin, J. Chiche, B. Zunino, M. Beneteau, O. Meynet, L.A. Pradelli,
J. Henschel, T. Steinfeldt, I.E. Gentle, G. Hacker, A non-death function of the S. Marchetti, A. Cornille, M. Carles, J.E. Ricci, GAPDH binds to active Akt, leading
mitochondrial apoptosis apparatus in immunity, EMBO J 38 (2019). to Bcl-xL increase and escape from caspase-independent cell death, Cell Death
[205] M.S. Soengas, P. Capodieci, D. Polsky, J. Mora, M. Esteller, X. Opitz-Araya, Differ. 20 (2013) 1043–1054.
R. McCombie, J.G. Herman, W.L. Gerald, Y.A. Lazebnik, C. Cordon-Cardo,

15

You might also like