Role of Bcl2 family proteins and caspases in the regulation of apoptosis

18
Role of Bcl-2 family proteins and caspases in the regulation of apoptosis Mohammad Shamsul Ola Mohd. Nawaz Haseeb Ahsan Received: 29 September 2010 / Accepted: 13 December 2010 / Published online: 6 January 2011 Ó Springer Science+Business Media, LLC. 2011 Abstract Apoptosis, or programmed cell death, plays a pivotal role in the elimination of unwanted, damaged, or infected cells in multicellular organisms and also in diverse biological processes, including development, cell differ- entiation, and proliferation. Apoptosis is a highly regulated form of cell death, and dysregulation of apoptosis results in pathological conditions including cancer, autoimmune and neurodegenerative diseases. The Bcl-2 family proteins are key regulators of apoptosis, which include both anti- and pro-apoptotic proteins, and a slight change in the dynamic balance of these proteins may result either in inhibition or promotion of cell death. Execution of apoptosis by various stimuli is initiated by activating either intrinsic or extrinsic pathways which lead to a series of downstream cascade of events, releasing of various apoptotic mediators from mitochondria and activation of caspases, important for the cell fate. In view of recent research advances about underlying mechanism of apoptosis, this review highlights the basics concept of apoptosis and its regulation by Bcl-2 family of protein. Furthermore, this review discusses the interplay of various apoptotic mediators and caspases to decide the fate of the cell. We expect that this review will add to the pool of basic information necessary to under- stand the mechanism of apoptosis which may implicate in designing better strategy to develop biomedical therapy to control apoptosis. Keywords Apoptosis Á Bcl-2 Á BH-3 only proteins Á Caspases Á Mitochondrial proteins Á Programmed cell death Abbreviations PCD Programmed cell death Bcl-2 B cell lymphoma-2 protein Bax Bcl-2 associated X protein Bid Bcl-2 interacting domain death agonist Bad Bcl-2 antagonist of cell death Bcl-xl Bcl-extra long Bim Bcl-2 interacting mediator of cell death Bik Bcl-2 interacting killer Bmf Bcl-2 modifying factor Boo Bcl-2 homolog of ovary Bcl-xs Bcl-extra short Bak Bcl-2 antagonistic killer Bok Bcl-2 related ovarian killer Apaf-1 Apoptosis protease-activating factor-1 Diablo Direct IAP binding Protein with low pI FADD Fas-associated death domain protein TNF-R Tumor necrosis factor receptor Fas-L Fas ligand HtrA High-temperature requirement IAP Inhibitor of apoptosis protein IMM Inner mitochondrial membrane Omi/HtrA2 Mammalian serine protease SMAC Second mitochondrial activator of caspase TNF-a Tumor Necrosis Factor alpha TRADD TNF-receptor-1 associated death domain protein VDAC Voltage-dependent anion channel Cyt c Cytochrome c PIDDosome p53-Inducible death domain containing protein complex DISC Death-inducing signaling complex M. S. Ola Á Mohd. Nawaz Department of Ophthalmology, College of Medicine, King Saud University, Riyadh 11411, KSA H. Ahsan (&) Department of Biochemistry, Faculty of Dentistry, Jamia Millia Islamia (A Central University), New Delhi 110025, India e-mail: [email protected] 123 Mol Cell Biochem (2011) 351:41–58 DOI 10.1007/s11010-010-0709-x

Transcript of Role of Bcl2 family proteins and caspases in the regulation of apoptosis

Role of Bcl-2 family proteins and caspases in the regulationof apoptosis

Mohammad Shamsul Ola • Mohd. Nawaz •

Haseeb Ahsan

Received: 29 September 2010 / Accepted: 13 December 2010 / Published online: 6 January 2011

� Springer Science+Business Media, LLC. 2011

Abstract Apoptosis, or programmed cell death, plays a

pivotal role in the elimination of unwanted, damaged, or

infected cells in multicellular organisms and also in diverse

biological processes, including development, cell differ-

entiation, and proliferation. Apoptosis is a highly regulated

form of cell death, and dysregulation of apoptosis results in

pathological conditions including cancer, autoimmune and

neurodegenerative diseases. The Bcl-2 family proteins are

key regulators of apoptosis, which include both anti- and

pro-apoptotic proteins, and a slight change in the dynamic

balance of these proteins may result either in inhibition or

promotion of cell death. Execution of apoptosis by various

stimuli is initiated by activating either intrinsic or extrinsic

pathways which lead to a series of downstream cascade of

events, releasing of various apoptotic mediators from

mitochondria and activation of caspases, important for the

cell fate. In view of recent research advances about

underlying mechanism of apoptosis, this review highlights

the basics concept of apoptosis and its regulation by Bcl-2

family of protein. Furthermore, this review discusses the

interplay of various apoptotic mediators and caspases to

decide the fate of the cell. We expect that this review will

add to the pool of basic information necessary to under-

stand the mechanism of apoptosis which may implicate in

designing better strategy to develop biomedical therapy to

control apoptosis.

Keywords Apoptosis � Bcl-2 � BH-3 only proteins �Caspases � Mitochondrial proteins � Programmed cell death

Abbreviations

PCD Programmed cell death

Bcl-2 B cell lymphoma-2 protein

Bax Bcl-2 associated X protein

Bid Bcl-2 interacting domain death agonist

Bad Bcl-2 antagonist of cell death

Bcl-xl Bcl-extra long

Bim Bcl-2 interacting mediator of cell death

Bik Bcl-2 interacting killer

Bmf Bcl-2 modifying factor

Boo Bcl-2 homolog of ovary

Bcl-xs Bcl-extra short

Bak Bcl-2 antagonistic killer

Bok Bcl-2 related ovarian killer

Apaf-1 Apoptosis protease-activating factor-1

Diablo Direct IAP binding Protein with low pI

FADD Fas-associated death domain protein

TNF-R Tumor necrosis factor receptor

Fas-L Fas ligand

HtrA High-temperature requirement

IAP Inhibitor of apoptosis protein

IMM Inner mitochondrial membrane

Omi/HtrA2 Mammalian serine protease

SMAC Second mitochondrial activator of caspase

TNF-a Tumor Necrosis Factor alpha

TRADD TNF-receptor-1 associated death domain

protein

VDAC Voltage-dependent anion channel

Cyt c Cytochrome c

PIDDosome p53-Inducible death domain containing

protein complex

DISC Death-inducing signaling complex

M. S. Ola � Mohd. Nawaz

Department of Ophthalmology, College of Medicine, King Saud

University, Riyadh 11411, KSA

H. Ahsan (&)

Department of Biochemistry, Faculty of Dentistry, Jamia Millia

Islamia (A Central University), New Delhi 110025, India

e-mail: [email protected]

123

Mol Cell Biochem (2011) 351:41–58

DOI 10.1007/s11010-010-0709-x

Introduction

Apoptosis is an ancient Greek word which means ‘‘the

falling of leaves from a tree’’. John Kerr first introduced the

name which refers to the morphological feature of forma-

tion of ‘‘apoptotic bodies’’ from a cell [1]. Apoptosis, also

known as programmed cell death (PCD), plays an integral

role in a variety of biological events including morpho-

genesis, tissue homeostasis, aging, and removal of

unwanted harmful cells [2]. Changes typical for apoptosis

include condensation of the nuclei, DNA fragmentation,

chromatin condensation, generation of envoluted mem-

brane segments, cellular shrinkage, and disintegration of

mitochondria [3, 4]. Apoptosis can be induced by a variety

of physiological and pathophysiological stimuli, such as

specific receptor molecules-CD95 [5, 6], tumor necrosis

factor (TNFa) [7], growth factors [8], ultraviolet light [9,

10], irradiation [11], heat shock [12], cytotoxic drugs [13,

14], oxidative stress [15], ceramide treatment [16], and

bacteria [17]. Dysfunction of apoptotic pathways result in

various pathological conditions including cancer, autoim-

mune, and neurodegenerative diseases [18–21].

A significant progress has been made toward under-

standing the mechanism of apoptosis and other apoptosis

related dysfunction, but it needs more intensive research to

better understand the whole cascade of events involved in

apoptosis and also the mechanism of their regulatory

pathways [22–24]. The nematode, Caenorhabditis elegans

(C. elegans), has been widely used to elucidate the

molecular pathways implicated in the regulation of PCD

[25–27]. For elucidating the underlying genetics and bio-

chemical pathways in PCD, three scientists, namely,

S. Brenner, J. Sulston, and R. Horvitz were awarded the

Noble Prize in medicine (2002).

The Bcl-2 and apoptosis

The underlying mechanism and genetics of apoptosis

began a new approach of research after the landmark dis-

covery by Horvitz, 1999 [28]. The protein expression of

two C. elegans cell death genes (CED-3 and CED-4) is

necessary for PCD during development, and CED-9 is the

functional homolog to mammalian protein Bcl-2, which is

shown to correct the phenotype of C. elegans by its anti-

apoptotic function [29, 30]. CED-9 was found to act

upstream of aspartate-directed cysteine protease (CED-3

and CED-4) known as caspases [31]. Previously, it was

revealed that proteins that are encoded by the mutant genes

discovered in C. elegans shared homology with mamma-

lian proteins, including Bcl-2 [32].

The role of Bcl-2 (B cell lymphoma 2), a founder

member of the Bcl-2 family of apoptosis regulator proteins,

has been elucidated in tumor development by dysfunction

in apoptotic pathways [20, 30]. In mammals, there is an

extrinsic apoptotic pathway that involves death receptors

and an intrinsic apoptotic pathway that involves the mito-

chondria. Mammalian mitochondria control apoptosis by

releasing many apoptogenic mediators from intermem-

brane space [33] and that further leads to a cascade of

reaction with the assistance of caspases, proteases that

cleave key cellular proteins [34, 35]. A major difference

between apoptosis in worm and mammals is the greater

role of mitochondria in the mammals but not in the worm.

Bcl-2 expression specifically blocks the morphologic fea-

tures of apoptosis, including the plasma membrane bleb-

bing, DNA cleavage, and nuclear condensation. Bcl-2

plays a role in cell survival and also inhibits cell death

induced by various stimuli such as chemotherapeutic

agents, ethanol and heat shock, indicating Bcl-2 as a neg-

ative regulator of cell death. The anti-death role is also

demonstrated in vivo by generation of mice lacking the

bcl-2 gene, which shows a variety of abnormalities

including excessive cell death.

Members and structural classification of Bcl-2 proteins

On the basis of various structural and functional charac-

teristics, the Bcl-2 family is divided into anti-apoptotic and

pro-apoptotic proteins [36–38]. All of the members of Bcl-

2 family have one or more homology domains labeled as

Bcl-2 homology (BH1, -2, -3, and -4) which are important

for heterodimeric interaction among members of the Bcl-2

family [27] (Fig. 1). Anti-apoptotic Bcl-2 family multido-

main proteins include Bcl-2, Bcl-xL (Bcl-extra long), A1,

Bcl-w, and Boo (Bcl-2 homolog of ovary) contain BH-

(1-4) domains. The Bcl-xL and Bcl-2 have a carboxy-ter-

minal hydrophobic transmembrane tail domain which helps

in localization of protein in outer mitochondrial membrane

(OMM) with the exception that Bcl-2 also resides in the

nuclear and endoplasmic reticulum membrane and is

translocated to OMM upon an apoptotic signal. Myeloid

cell leukemia factor-1 (Mcl-1) is the only antiapoptotic

Bcl-2 protein with three BH domains (BH-1, -2, and -3).

Pro-apoptotic Bcl-2 family of proteins are divided into

two subgroups according to the number of BH domains

(e.g., Bax, Bak, and Box) or those proteins that have only

the BH3 domain (e.g., Bid, Bim, and Bad) with the

exception of Bcl-xS which has only BH3 and BH4 domains

(Fig. 1). There are eight members of BH3 only protein,

which include hara-kiri (Hrk), BH3 interacting domain

death agonist (Bid), Bcl-2 interacting mediator of cell

death (Bim), Bcl-2 modifying factor (Bmf), p53, promoter-

upregulated modulator of apoptosis (Puma), Noxa (named

for ‘‘damage’’), Bcl-2 antagonist of cell death (Bad), and

42 Mol Cell Biochem (2011) 351:41–58

123

Bcl-2 interacting killer (Bik). Bid, Bad, and Bim have

cytosolic location with respect to mitochondria. Following

death signal, BH3 domain-only proteins also known as

minimal death domain can neutralize or depress antiapo-

ptotic Bcl-2 proteins allowing pro-apoptotic proteins such

as Bax/Bak-like proteins to induce apoptosis [39, 40].

A restricted subset of these BH3 proteins undergoes

post-translation modification such as cleavage, phosphor-

ylation, transcriptional upregulation, and dissociation. Such

modifications result in their activation and translocation to

outer mitochondrial membrane (OMM) which helps it to

interact with multidomain pro-apototic members, Bax/Bak-

like proteins, leading to their oligomerization and forma-

tion of pore. Such pore may further lead to release of other

pro-apoptotic protein on the intermembrane space (IMS)

that enhances the processes of apoptosis [41]. Thus the

regulation of expression and function of BH3 domain-only

proteins is critical in mediating cell death and under-

standing their expression and function necessary for cyto-

protective strategies. In general, the relative ratio of pro-

survival (Bcl-2) and pro-apoptotic (Bax and BH3-only)

proteins seems to determine the cell sensitivity or resis-

tance to the apoptotic stimuli.

Mechanism of action and role of Bcl-2 family members

in the regulation of mitochondrial apoptosis

Antiapoptotic or pro-survival proteins

It is now becoming evident that almost every nucleated cell

requires protection by at least one of the member Bcl-2

homolog, and that regulates homeostasis in mammalian cell

and tissues. Over-expression of Bcl-2 protein in hemato-

poietic lineages yields excess T, B, and myeloid cells and

enhance the survival of these progenitor cells that are hard to

avoid cytotoxic insults [29, 42]. Alternatively, inactivation

of the Bcl-2 gene increases apoptosis in specific cell types,

presumably because the concentrations of other Bcl-2

homolog are too low to compensate. Bcl-2 itself is required

for the survival of kidney, mature lymphocytes and mela-

nocyte stem cells [43], Bcl-w for sperm progenitors in adult

mice [44], an A1 gene for neutrophils [45], Mcl-1 for zygote

implantation [46], and Bcl-xL for erythroid and neuronal

cells [47]. The hydrophobic carboxy-terminal domain of

these antiapoptotic proteins help to anchor themselves to the

cytoplasmic face of three intracellular membranes: the outer

mitochondrial membrane, the nuclear envelope, and the

endoplasmic reticulum (ER). Bcl-2 is an integral membrane

protein, whereas Bcl-xL and Bcl-w only become tightly

associated with the membrane after cytotoxic signals which

induce conformational changes to protect cell. The antiap-

optotic member of Bcl-2 family, including Bcl-2 itself,

decreases apoptosis mainly by preventing mitochondrial

outer membrane permeabilization (MOMP) either by neu-

tralizing the activity of both pro-apoptotic members (BH3

only protein and multidomain members) [48]. If Bcl-2 or

Bcl-xl cannot exert their antiapoptotic property, Bax oligo-

merization takes place and apoptosis is promoted.

Pro-apoptotic or anti-survival protein

It includes both multi-domain and BH3-only protein

members.

Multi-domain pro-apoptotic member

Antiapoptotic members such as Bcl-2 interact with Bak and

Bax and inhibit their oligomerization and/or bind to BH3

only proteins to block apoptosis [48]. Bak interacts with

two mitochondrial fusion proteins (Mfn1 and Mfn2). After

receiving apoptotic signal, Bak dissociates from Mfn2 to

become associated with Mfn1. It has been shown that

reconstitution of Bak into Bax/Bak double knockout cells

Fig. 1 According to conserve

domain (BH) and its anti or pro-

apoptotic action, the Bcl-2

super-family can be divided into

three sub-families: Anti-

apoptotic or Pro-survival multi-

domain, pro-apoptotic multi-

domain and pro-apoptotic BH3

only. Few examples of protein

from each group of the family

are given in parenthesis. TM

(transmembrane domain)

Mol Cell Biochem (2011) 351:41–58 43

123

restores mitochondrial fragmentation and Bak in associa-

tion with Bax permeabilize the outer membrane of mito-

chondria rendering apoptotic cascade [49]. Bak may

contribute to early mitochondrial fragmentation while Bax

is probably more important for subsequent pores develop-

ment and degeneration in the outer membrane [49]. A

number of studies have shown that insertion of activated

oligomerized Bax and/or Bak into the OMM exerts stress

on the membrane, leading to supramolecular pores that

include lipids (lipidic pores) in the OMM [50–52].

In healthy cells, Bak is inactive in the OMM through its

association with Voltage-dependent Anion channel

(VDAC-2) [48], whereas Bax is dormant in the cytosol

through interactions with several proteins, including

Ku-70, 14-3-3, and the humanin peptide. Many apoptotic

signals can trigger BH3-only-protein-dependent transloca-

tion of Bax, followed by its insertion into the OMM and the

formation of Bak or Bax homo-oligomers. Several models

have been proposed for how Bcl-2 anti-apoptotic proteins

antagonize the functions of Bax, Bak, and BH3-only pro-

teins. Activation of Bax and Bak leads to a conformational

change that exposes the N-terminus of the proteins, which is

otherwise hidden in the inactive state [53]. After Bak and

Bax form homo-oligomers, they participate in forming pores

in OMM and cause permeabilization leading to the release of

the contents of the mitochondrial intermembrane space

(IMS) including Smac and cytochrome c into the cytosol

[54, 55]. The released mitochondrial protein then activates

caspases, which causes a series of cascade reactions that

cleave essential proteins complement throughout the cell.

Two models of Bax and Bak activation, exist the direct

and indirect models. According to the direct model various

BH3 pro-apoptotic proteins called as activator proteins (Bid,

Bim, Puma, and p53) directly interact and induce confor-

mational changes in Bak and Bax [56]. Members of anti-

apoptotic proteins bind and sequester such activator, and

also bind to activated Bak and Bax proteins that might be

present and thus prevent apoptosis. BH3-only proteins are

further divided into activator and sensitizer categories [57].

Sensitizers bind to anti-apoptotic proteins and cause the

release of activator BH3-only proteins leading to activation

of Bak and Bax. However, it is likely that additional factors

other than Bid and Bim can act as activators such as Puma,

p53, and heat as activators of Bax and Bak [58–60], and

possibly others. Gavathiotis and colleagues have demon-

strated that stabilized alpha-helix of Bcl-2 domains (SA-

HBs) directly initiates Bax-mediated mitochondrial

apoptosis [61].

According to the indirect model, anti-apoptotic proteins

always bind to Bak and Bax and prevent their activation,

whereas in response to death signal, BH3-only proteins

bind to antiapoptotic proteins causing its release and ini-

tiating death signal through activation of Bax and Bak [62].

The sequestered forms of Bak and Bax are those fractions

of the total Bak and Bax population that are already

activated either spontaneously, or by other unspecified

means [63]. Bax and Bak can also alternatively interact

with proteins that remove the requirement for sequestra-

tion by anti-apoptotic proteins [64]. Activated population

of Bak and Bax are needed to kill, and that should be

sequestered by anti-apoptotic proteins to maintain survival

of the cell. Activated Bak and Bax are responsible for the

permeabilization of membranes. The activation is

achieved, either by interacting with activator proteins or

through some agents, while the anti-apoptotic proteins

inhibit death by sequestering activated Bak and Bax or

activator proteins.

BH3-only pro-apoptotic members

BH3-only pro-apoptotic proteins are involved in main-

taining the functionality and cellular integrity of the cell in

which Bim acts as sensor for cytoskeleton integrity, Bid

acts as sensor for death domain receptor signaling, and Bad

acts as an inhibitor for growth factors withdrawal. Indi-

vidual BH3-only proteins are normally controlled by

diverse mechanisms.

Cleavage of Bid Activation of cell surface receptors such

as tumor necrosis factor receptor (TNF-R) activates cas-

pase-8, which leads to the cleavage of Bid, and then

truncated Bid (active Bid or tBid) translocates from cytosol

to OMM and induces cytochrome c release [65–67]. Oh

and his group have demonstrated that at nanomolar con-

centrations of a synthetic Bid activates Bax almost as

efficiently as tBid itself thus tBid engages Bax to trigger its

pro-apoptotic activity [68]. It is now well established that

tBid leads to oligomerization of Bax and Bak, which forms

pores in OMM which in turn permits the release of IMS

proteins into the cytosol [54, 69]. During apoptosis, Bid

induces the mobilization of cyt c by remodeling mito-

chondrial cristae by interacting Bid with cardiolipin [70,

71]. Addition of tBid to permeabilized cells leads to

hydrolysis of cardiolipin molecules thereby decreasing the

cardiolipin levels [72, 73]. But tBid alone does not cause

mitochondrial outer membrane permeabilization [74].

Phosphorylation of Bad In the absence of various sur-

vival factors, Bad is activated by dephosphorylation [75].

The BH3 domain of Bad binds and inactivates Bcl-2 and/or

Bcl-xL at the OMM, thereby promoting cell death. Con-

versely, in the presence of survival factors, Bad is phos-

phorylated, making it to dissociate from Bcl-2 and/or Bcl-

xl, permitting, and survival promotion. There are several

phosphatases that dephosphorylate Bad in vitro [75, 76].

44 Mol Cell Biochem (2011) 351:41–58

123

However, a report by Datta and his group showed that Akt

phosphorylates Bad, both in vitro and in vivo, and blocks

the Bad-induced death of primary neurons [77]. Further-

more, several studies elucidate the importance of Bad

phosphorylation and underlying mechanism of cell death

[78–80].

Dissociation of Bim and Bmf In response to several

apoptotic stimuli, such as detachment of adherent cells

from their substratum (anoikis) or ultraviolet irradiation,

Bmf is released from the myosin V motor complex,

translocates and binds to antiapoptotic family proteins,

such as Bcl-2, Bcl-xl, and Bcl-w, but does not interact with

the pro-apoptotic family protein, such as Bax, Bid, and Bad

[81]. However, c-Jun N-terminal protein kinase (JNK) that

causes phosphorylation of Bim is involved in ischemia-

induced neuronal apoptosis through activation and trans-

location of Bax [82, 83]. Bim is the main regulator of

hematopoietic homeostasis [84], essential for the elimina-

tion of autoreactive lymphocytes [85] and plays vital role

in neuronal death [86]. Recently, it has been shown that

Bim displaces Bcl-xl in the mitochondria and promotes

Bax translocation during intrinsic pathways assisted by

TNFa [87] or UV-induced [88] apoptosis. Histone deace-

tylase (HDAC) inhibitors increase ionizing radiation-

induced apoptosis in several cancer cells via activation of

Bmf transcription [89, 90]. In TGF beta-induced cell death,

there is upregulation of Bmf and Bim, and thus, inhibition

of the TGFb provides an important therapeutic and pro-

tection of cells from apoptosis [91].

Transcriptional regulation of BH3-only proteins BH3-

only proteins are also transcriptionally regulated as in DNA

damage-induced apoptosis which requires synthesis of new

protein. Transcription of Noxa and Puma, the BH3-only

group members, is induced by p53 [92–94]. Apoptosis in

fibroblast in response to DNA damage is decreased in mice

knocked out in Noxa/puma whereas mutations of the BH3

domain suppress the pro-apoptotic activity of Noxa [94],

and Noxa-induced apoptosis is inhibited by Bcl-xl and

Bcl-2 [95]. Recently, it has been reported that Puma

expressed independent of p53 regulation [96, 97] and ini-

tiates apoptosis by dissociating Bcl-xl and Bax, promoting

Bax multimerization and mitochondrial translocation

[9, 96]. Puma rapidly induces apoptosis in cells lacking not

only the BH3-only proteins but in the absence of Bid and

Bim [98]. Puma induced apoptosis is associated by

regeneration of superoxide and H2O2 which is Bax-

dependent and that can be confirmed by the presence of

antioxidants that prevent Puma-dependent apoptosis [99].

Whereas Bax inactivation confers a resistance to Puma-

dependent apoptosis [100].

Apoptotic pathways

Apoptosis or PCD has conserved genetic and biochemical

pathways [27, 101]. In vertebrates, caspase-dependent

apoptosis occurs through two main interconnected path-

ways which are intrinsic and extrinsic pathways [102]. The

intrinsic pathway or intracellular path is mediated by Bcl-2

family, whereas the death receptor or extrinsic pathway is

activated by signal from other cells [7, 103, 104].

Intrinsic pathway

The intrinsic pathways also known as mitochondrial path-

ways or stress pathways are activated by a diverse array of

death stress, genomic stress, metabolic stress, presence of

unfolded proteins, and other stimuli that lead to perme-

abilization of OMM and release of apoptotic proteins into

the cytosol (Fig. 2). Several of these proteins including

cytochrome c (cyt c) initiate or regulate caspase activation.

The cyt c plays the main role in this pathway which is

activated after its interaction with apoptotic protease-acti-

vating factor (Apaf1) and deoxyadenosine triphosphate

(dATP) to form apoptosome [102, 105]. The apoptosome

creates a platform to bring together molecules of the ini-

tiator caspase of the intrinsic pathway. Progression through

the pathway usually leads to activation of caspase-9,

enabling their auto-activation. Caspase-8 also has a major

role to play by activating the Bid leading to the formation

of activated truncated Bid (tBid), which translocates to

mitochondria and releases cyt c. The activated caspase-8

then cleaves procaspase-3, giving activated caspase-3,

which acts as an executioner, by cleaving multiple of other

substrate within the cells [34, 106].

Extrinsic pathway

The extrinsic pathway involves the association of receptor-

mediated transmembrane death receptor (FAS and TNF-a)

and its extracellular ligand (FAS-L and TNF-aL) [107–

109]. For association the receptor trimerizes and death

adapter molecules are recruited on the cytoplasmic side of

the mitochondrial membrane. Fas recruits Fas-associated

death domain protein (FADD) whereas TNF-R recruits

TNF-R1-associated death domain protein (TRADD) which

again recruits FADD [7]. Such association leads to for-

mation of death-inducing signaling complex (DISC) con-

sisting of a complex of receptor, its ligand, the initiator

procaspase-8 (or procaspase-10), and some other regulators

and co-factor [110]. The complex helps in recruiting more

of procaspase-8 (or procaspase-10) and enables their

autoactivation (Fig. 2).

Mol Cell Biochem (2011) 351:41–58 45

123

Fig. 2 The underlying mechanisms in the execution of apoptotic

pathways. The Apoptotic stimulus received are communicated to the

cell by extrinsic or intrinsic pathways. The extrinsic pathway involves

the engagement of various transmembrane death receptors such as

FAS and TNF-a and their corresponding extracellular ligands (FASL,

TNF-aL). Fas recruits, Fas-associated death domain protein (FADD)

whereas TNF-R recruits TNF-R1 associated death domain protein

(TRADD) which again recruits FADD. Such association leads to

formation of Death-inducing signaling complex (DISC) which acts as

platform for many co-factors and regulator for cleavage and

activation of procaspase-8 (or 10), an initiator caspase of the extrinsic

pathway. The activated caspase-8 (or 10) leads to their own

autoactivation which further activate the caspase-3 and caspase-7,

an effectors caspase. The intrinsic pathway involves the mitochondria

and after receiving the stimulus, the pro-apototic BH3-only members

(Bid, bad, bim, etc.) bind and neutralize the antiapoptotic proteins

(Bcl-2, Bcl-xL, etc.). This leads to oligomerization of multi-domain

pro-apoptotic member (Bax/Bak) present on mitochondrial membrane

surface whose activation leads to permeabilization and formation of

pores or VDAC in outer mitochondrial membrane (OMM), releasing

various apoptotic mediators (HtrA2/Omi, Smac/DIABLO, Cyt c, endo

G and AIF). The release of Cyt c in cytosol causes the association of

Apaf-1 and ATP/dATP to form apoptosome, the initiator caspase of

the intrinsic pathway, caspase-9, enabling their autoactivation. This in

turn activates the series of downstream of cascade to activate the

caspase-3 and caspase-9, an effectors caspase involved in cell death.

The role of IAP is to act a guardian inside a cell to defend against the

mediator of apoptosis (HtrA2/Omi, Smac/DIABLO) by binding to

Caspase 3/7/9 where as Endo G and AIF acts in a caspase independent

manner by causing chromatin condensation and fragmentation. The

pro-apototic BH3-only members, Bid, is also activated by caspase-8,

making it an important converging point of both pathways. The

activated truncated (t)Bid further activates downstream targets, such

as BAX and BAK

46 Mol Cell Biochem (2011) 351:41–58

123

Apoptotic mediators

Permeabilization of the outer mitochondrial membrane

allows the leakage of at least five apoptotic mediators

(apoptogenic proteins) from the mitochondrial intermem-

brane space, such as cyt c, second mitochondrial activator

of caspases/Direct IAP protein with low pI (Smac/DIA-

BLO), HtrA2/Omi, apoptosis-inducing factors (AIF), and

endonuclease G [32, 37]. These proteins induce apoptosis

in different ways. Smac/DIABLO and HtrA2/Omi suppress

the ability of IAPs (inhibitors of apoptosis proteins) to

inhibit caspases. Endonuclease G and AIF are involved in

DNA fragmentation, and AIF is also involved in chromatin

condensation (Fig. 2). The release of these apoptotic

mediators from mitochondria is known to be regulated by

Bcl-2 family of proteins [55, 111]. However, caspase-

independent mitochondrial cell death results from loss of

respiration and not from the release of cytotoxic apoptotic

mediators [112].

Release of cytochrome c

Cyt c is a water soluble 13 kDa protein encoded by nuclear

gene that is translated in cytosol to be finally imported into

mitochondria. It normally resides in the spaces within

cristae of the inner mitochondrial membrane (IMM) and at

narrow cristae junctions [113]. The role of cyt c in the

intrinsic pathways in mammalian cells is well known.

Addition of dATP to cytosolic extract induces caspase

activity [114] and depletion of cyt c in cell extract inhibits

its apoptotic potential and also microinjection of cyt c in

various cell types enhances the apoptotic pathways [115,

116]. Thus, cyt c is the main mediator of apoptosis [117,

118], and the release of cyt c occurs due to DNA damage

[119, 120]. Over-expression of Bcl-2 blocks the release of

cyt c from mitochondria and inhibits the initiation of

apoptosis [121].

Release of endonuclease G (EndoG)

EndoG is a 30 kDa nuclease protein located in mitochon-

drial intermembrane space [122, 123]. The release of

EndoG after apoptotic signal leads to DNA fragmentation

as found in inhibitor of caspase-activated DNase (ICAD)-

deficient cells after induction of apoptosis by TNF treat-

ment and UV-irradiation [123, 124]. Once released, EndoG

participates in DNA fragmentation but without assistance

of caspases [125–127].

Release of apoptosis-inducing factors (AIF)

Apoptosis-inducing factor (AIF), which resembles bacterial

oxidoreductase, is a 57 kDa flavoprotein present in the

mitochondrial intermembrane space [128]. Upon induction

of apoptosis, AIF translocates from the mitochondria to the

nucleus and causes DNA fragmentation and chromatin

condensation [128]. These effects are independent of its

oxidoreductase and caspases activity [129, 130]. Disruption

of AIF in mice prevents normal apoptosis necessary for the

activation of embryoid bodies in the embryo [131]. In

addition, AIF is required for specific cell death pathways

including lethal responses to excitotoxins such as gluta-

mate and N-methyl-D-aspartate (NMDA), DNA-alkylating

agents, hypoxia–ischemia, or growth factor deprivation

[132]. Recently, Schulthess and coworkers have shown a

protective role of AIF on b-cell turnover and the loss of

AIF increases b-cell apoptosis. AIF is essential for main-

taining b-cell mass and oxidative stress response [133].

The mechanism of AIF-induced large-scale chromatin

condensation and DNA fragmentation is still not clear

[128, 131, 134–136]. More recently, the role of AIF in

12/15-lipoxygenase (LOX)-dependent organelle damage

pathway has been reported showing that AIF and 12/15-

LOX are important mediators in a common cell death

pathway in stroke-induced brain damage [137].

Release of Smac/DIABLO

SMAC (Second Mitochondrial Activator of Caspases) or

DIABLO (Direct IAP Binding Protein with Low pI) are a

25 kDa, pro-apoptotic protein released from the inter-

membrane space that neutralizes the inhibitory activity of

IAP leading to activation of caspases and apoptosis [138–

140]. During reovirus-induced apoptosis, Smac/DIABLO

are released that decreases the level of IAP’s and thus

activates apoptosis [139]. Sphingosine 1-phosphate inhibits

the release of Smac/DIABLO from mitochondria and

antagonizes apoptosis of human leukemia cells. Smac/

DIABLO is involved in many cancer manifestation and

progression [141] such as cervical cancer [142], colon

cancer [143], and hepatocellular carcinoma [144].

Release of Omi/HtrA2

The mammalian serine protease Omi/HtrA2 (high-tem-

perature requirement) is a 49 kD protein, homologous to

the bacterial endoprotease also known as DegP [145].

Unlike Smac/DIABLO, the pro-apoptotic activity of Omi/

HtrA2 involves both IAP binding and serine protease

activity. Omi/HtrA2 has a dual function, when residing

inside the mitochondria it promotes cell survival, but when

released into the cytoplasm it participates in both caspase-

dependent and -independent cell death. It prevents the

IAP’s action via amino-terminal reaper-related motif which

induces caspase activity [99, 146–148]. It also mediates

caspase-independent cell death through its own protease

Mol Cell Biochem (2011) 351:41–58 47

123

activity, by the fact that simultaneous deletion of the other

IAP binding protein, Smac/DIABLO, does not alter the

phenotype of Omi/HtrA2 knockout mice or cells derived

from them [149]. The caspase-independent role of Omi/

HtrA2 in apoptosis is evident in human cardiac-specific

inhibitor of cell cycle protein, Thanatos-Associated Protein

5 (THAP5) that are cleaved by pro-apoptotic Omi/HtrA2

during cardiomyocytic cell death [150]. The role of Omi/

HtrA2 in colon cancer [151] and oxidative stress of pig-

ment epithelial cell [152] further supports its application of

mediating apoptosis in cells. The role of Omi/HtrA2 in

promoting cell death by binding and degrading ped/pea-15,

an anti-apoptotic protein, establishes the pro-apoptotic

effect of Omi/HtrA2 [153] and its role in the apoptosis of

prostate cancer cell, PC-3 [154]. However, a contradictory

role of mitochondrial Omi/HtrA2 has been reported. In

response to some extracellular inducers of mitochondrial

stress, Omi/HtrA2 stabilizes mitochondrial membrane

potential and inhibits mitochondrial superoxide generation

and hence controls apoptosis [155].

Regulation of apoptosis by IAPs

Inhibitor of apoptosis (IAPs) is a family of antiapoptotic

proteins that associate with caspases in response to diverse

stimuli. IAP has been discovered both in invertebrate and

vertebrates. So far, eight human IAP homologs have been

identified which includes X-chromosome-linked IAP

(XIAP, also known as hILP, MIHA, or BIRC4), survivin

(also known as TIAP or BIRC5), cellular IAP1 (c-IAP1, also

known as HIAP2, MIHB or BIRC2), c-IAP2 (also known as

HIAP1, MIHC, or BIRC3), neuronal apoptosis inhibitory

protein (also known as BIRC1), IAP-like protein 2 (also

known as BIRC8, or Ts-IAP), apollon (also known as Bruce

or BIRC6), melanoma IAP (ML-IAP, also known as KIAP,

livin, or BIRC7) [156, 157]. The best characterized IAPs

such as XIAP, c-IAP1, and c-IAP2, bind caspase-3, caspase-

7, and caspase-9, thereby inhibiting their activation and

preventing apoptosis (Table 1) [156, 158–160]. Also, cIAP1

and cIAP2 have been shown to bind caspases, although how

they inhibit apoptosis at the molecular level is not com-

pletely understood [161]. Direct inhibition of caspase

activity by c-IAPs is an important means of regulation in

order to protect cell death. Activity of XIAP is blocked by

binding to Omi/Htr2A and Smac/DIABLO proteins released

from mitochondria after pro-apoptotic stimuli. Thus, Smac/

DIABLO is a negative regulator of IAPs [140, 162]. Simi-

larly, Omi/HtrA2 also inhibits the XIAP through a reaper-

like motif [148] and has a prognostic significance in hepa-

tocellular [144] and renal cell carcinoma [163]. Inhibition of

apoptosis increases the survival rate of cancer cells and

facilitates their escape from cytotoxic therapies and immune

surveillance [144, 164, 165].

Caspases and underlying mechanism in controlling

cellular apoptosis

Caspases are not only involved in the process of apoptosis

but also needed for the development and maturation of

cytokines leading to cell growth and differentiation [166].

Apoptotic cell death is dependent on a family of aspartate-

specific cysteine proteases (caspases) that cleave certain

vital structural proteins (e.g., lamins, gelsolin) and pro-

teolytically activate latent enzymes (e.g., nucleases) that

contribute to cell death. These enzymes exist in most cells

as inactive precursors (procaspases) that are converted into

their active forms by proteolytic cleavage at internal

aspartic acid residues, which separates the caspase into

small and large subunits [167] and then they become

activated by autoproteolysis.

Through genetic analysis of cell-death defective (CED)

mutants, it was found that the product of the aspartate-

directed cysteine protease (CED-3) is required for all

developmental-related programmed cell deaths in the

worm [168]. In this multicellular model, various genes that

encode for proteins essential for the regulation and exe-

cution of apoptosis were identified and their mammalian

homologs were described [33, 76]. For example, ced-9 and

Bcl-2 encode proteins that inhibit apoptosis, Apaf-1 and

ced-4 encode an adaptor protein that permits the interaction

between initiator proteins ced-3 and caspase-9. Ced-3

encodes a protein homologous to caspase-9 which is

responsible for the apoptotic initiation process. Fifteen

mammalian caspases have been identified, with caspase-11

and 12 identified only in mouse as shown in Table 1.

All caspases are synthesized as zymogens sharing a com-

mon domain structure consisting of a large (p10) and a small

(p20) catalytic subunit. Among two fundamentally different

groups, the first group, termed as initiator caspases, is char-

acterized by a long prodomain which provides a protein–

protein interaction platform [167, 169]. Prodomains allow the

recruitment of procaspases into an activating protein com-

plex. Long prodomain caspases are caspase-1, -2, -4, -5, -9,

-11, and -12 with an N-terminal caspase-activating recruit-

ment domain (CARD), and caspase-8 and -10 with an

N-terminal death effector domain (DED) (Table 1). In con-

trast to the initiator caspases, the second group consisting of

executioner caspases-3, -6, and -7 lack the large N-terminal

non-enzymatic domain and they are responsible for the

majority of cellular destruction during apoptosis [170]. When

procaspase-8 or -10 recruited to ligate death receptors by Fas-

Associated Death Domain (FADD), they undergo autocatal-

ysis, releasing the p10 and p20 subunits that form the active

(tetrameric) enzyme. Caspase-9 is activated in the presence of

ATP and cyt c by an allosteric change on a heptameric scaf-

fold of apoptotic protease-activating factor 1 (Apaf1) proteins

termed as apoptosome. Caspase-9 processing occurs

48 Mol Cell Biochem (2011) 351:41–58

123

secondary to caspase-3 activation during Smac-induced

apoptosis. In a heat shock-induced death, Caspase-2 induces

apoptosis via cleavage of Bid [171].

Interdependence of caspases and Bcl-2 in the regulation

of apoptosis

Numerous studies suggest that many apoptotic signaling

pathways converge at the mitochondria, where signals are

processed through a series of molecular events culminating

in the release of potent death factors that trigger either

through the extrinsic or the intrinsic pathway. Basically,

the release of the pro-apoptotic proteins from the inter-

membrane space triggers apoptosis, in a caspase-dependent

(through cyt c, Omi/HtrA2, and SMAC) as well as in a

caspase-independent form (through AIF and Endo G).

In mammalian cells, caspases-9, -8, and -2 rely on the

formation of apoptosome (Apaf-1), death-inducing signal-

ing complex (DISC), and PIDDosome, respectively, for

activation of apoptotic signals. Apoptosome is composed of

seven molecules (heptamer) of Apaf-1 bound to cyt c in the

presence of ATP/dATP, Fig. 2. DISC is assembled follow-

ing binding of death ligand to its receptor and contains

FADD and caspase-8 (or –10) whereas PIDDosome contains

at least three components, PIDD, RAIDD, and caspase-2.

Apoptosome complex

Apoptosome is a multimeric protein complex that mediates

activation of an initiator caspase at the onset of apoptosis.

Biochemical and structural investigations revealed

insights into the assembly and function of the various

apoptosomes from fruit fly (Drosophilia melanogaster),

worm (C. elgans) and mammals [172, 173]. The assembly

of the mammalian apoptosome which is responsible for the

activation of caspase-9 requires the binding of Apaf-1, cyt

c, and ATP/dATP. The apoptosome is an oligomeric sig-

naling platform that has a core of seven apoptotic Apaf1.

Each Apaf-1 monomer contains an N-terminal caspase

recruitment domain (CARD), a nucleotide-binding and

oligomerization domain (NOD), and a string of WD40

(tryptophan-aspartic acid) repeats at the C terminus

(Fig. 2). The WD40 repeats are thought to be the site of cyt

c binding [102, 174].

Death-inducing signaling complex (DISC)

DISC consist of complex of the death receptor (FAS), the

adaptor FAS-associated death domain protein (FADD), the

initiator caspase procaspase-8 (or procaspase-10), and

possibly other co-factors and regulators [175]. Kischkel

and coworkers reported the formation of a protein complex

in the dying cell and named it as DISC. Upon receiving the

death signal the activated death ligands homo-trimerizes,

which in turn, induces oligomerization of the Fas death

receptors (Fig. 2) [176, 177]. Dimerization of caspase-8 is

a crucial factor for activation and suggests that DISC may

facilitate the activation of caspase-8 through dimerization

[178]. The DISC creates a platform that brings together

molecules of the initiator caspase-8 (or caspase-10) and

other co-factors and regulators for execution of extrinsic

Table 1 Subfamily of caspase showing their functional role, functional domain and corresponding specificity to member of Inhibitor of

Apoptosis (IAP’s)

Sub-family Caspase Mammals Funtional domain Specificity to IAP’s

I- Caspase-2 Human CARD –

Apoptosis activator –

Caspase-8 Human DED XIAP, c-IAP1, c-IAP2, Ts-IAP, Ts-IAP

Caspase-9 Human DED –

Caspase-10 Human CARD –

II- Caspase-3 Human NED NAIP, XIAP, c-IAP1, c-IAP2, Survivin, livin

Apoptosis executioner Caspase-6 Human NED –

Caspase-7 Human NED NAIP, XIAP, c-IAP1, c-IAP2, Survivin, livin

III- Caspase-1 Human CARD –

Inflammatory mediator Caspase-4 Human CARD –

Caspase-5 Human CARD –

Caspase-11 Mouse CARD –

Caspase-12 Mouse CARD –

Caspase-13 Human CARD –

Caspase-14 Human CARD –

Caspase-14 Human CARD –

Caspase-Activating Recruitment domain (CARD), Death Effector Domain (DED), and Non-Enzymatic Domain (NED)

Mol Cell Biochem (2011) 351:41–58 49

123

apoptotic pathway which finally leads to their auto-acti-

vation [179, 180].

p53-Inducible death domain containing protein

complex (PIDDosome)

The PIDDosome complex under physiological conditions

contains PIDD and associate with the activation of another

initiator caspase, caspase-2 [181]. Although, the role of the

PIDDosome in apoptosis remains controversial, its

expression is inducible upon DNA damage [173, 181, 182].

However, PIDD-deficient mice undergo apoptosis not only

in response to DNA damage, but also in response to various

p53-independent stress signals and to death receptor

engagement. In the absence of PIDD, both caspase-2 pro-

cessing and activation occur in response to DNA damage

indicating that PIDD does not play an essential role for all

p53-mediated or p53-independent apoptotic pathways

[182]. The role of caspase-2 in the mitochondrial pathway

is now widely accepted [183–185]. Thus, the initial stage

of DNA damage facilitated by p53-mediated apoptosis

occurs by a PIDD and caspase 2-dependent mechanism.

For events that are downstream of cyt c release, p53’s full

transcriptional regulatory functions are required [186].

Apoptosis and the caspase-independent pathway

There also exists a caspase-independent apoptotic pathway

that is associated to AIF [131, 134], endonuclease G [187],

as well as Omi/HtrA2 [150–152]. As already discussed

above in this review, cells after receiving the apoptotic

signals release the nuclear AIF molecule and endonuclease

G protein which are translocated to the nucleus causing a

large-scale chromatin condensation and DNA fragmenta-

tion independently of caspase activation. Similarly Omi/

HtrA2 mediates caspase-independent cell death through its

own protease activity as can be observed during apoptosis.

A specific Omi/HtrA2 inhibitor can stop degradation of

THAP5 protein (THAP family of proteins), which leads to

reduced cell death [150]. A few studies have focused the

role of Heat Shock Proteins (HSPs) that are either consti-

tutively expressed or expressed under variety of stresses

stimuli [188], have shown their role in apoptosis via cas-

pase-independent pathway [189]. Members of HSP protein

such as HSP27 and HSP70 participate in oncogenesis,

probably by interfering apoptotic pathways. First, they act

as chaperones and play a role in proteasome-mediated

degradation of apoptosis-regulatory proteins. Second, they

inhibit key effectors of the apoptotic machinery including

the apoptosome and apoptosis-inducing factor [190].

However recent work on caspase-independent apoptotic

pathway has lead to the discovery of various other

molecules which has a major role in the pathway. For

example, the work by Yuan and his group have shown that

Ste20-like protein kinase 3 respond to apoptosis of HeLa

cells to trigger the caspase-independent apoptotic pathway

[191]. Similarly activated analog of CY, 4-hydroperoxy-

cyclophosphamide (4-OOH-CY) is being used for the

therapy for hematological malignancies and autoimmune

disorders through caspase-independent T-cell apoptosis

[192].

Therapeutic implication in the control of apoptosis

Lack of the phenomena of apoptosis results in excessive

increase of cell number and has implications in autoim-

munity and tumorigenesis. On the other hand, excessive

apoptosis decreases the cell population, which is linked to

many neurodegenerative disorders such as Parkinson’s

disease, Alzheimer’s disease, Huntington’s disease, and

spinal muscular atrophy. Alzheimer’s disease is a complex

neurological disorder in which the beta-amyloid peptides

are formed in the brain. In this, the Bcl-2 is down-regulated

and Bax is up-regulated [193, 194]. With respect to the role

of Bcl-2 proteins and its role in Parkinson’s disease, it has

been anticipated that pro-apoptotic family members par-

ticipate in neuronal death in a variety of Parkinson’s dis-

ease models [195]. Numerous studies show that activation

of apoptosis has also been found to be involved in the

pathogenesis of other human diseases such as chronic heart

failure [196], diabetes [197, 198], and atherosclerosis

[199]. The increased expression of Bcl-2 in the vascular

endothelium inhibits the diabetes-induced degeneration of

retinal capillaries and superoxide generation [198, 200].

An imbalance among the Bcl-2 family of proteins, in

favor of the anti-apoptotic members, is a phenomenon that

naturally, and frequently occurs in cancer cells [201–203].

Over-expression of anti-apoptotic Bcl-2 or Bcl-xl probably

occurs in more than half of all cancers. Moreover, loss of

expression of Bax is also found in some colorectal cancers

and in hematopoietic malignancies, whereas the expression

of a highly apoptogenic variant of Bax (Baxw) is correlated

with an increased survival of patients with glioblastoma

multiforme, an aggressive form of brain tumors [204]. The

defects may arise from the fact that neoplastic cells are

under strong selective pressure to stabilize their mito-

chondrial permeability, even if they harbor alterations in

the p53 tumor suppressor pathway. c-Myc, for instance,

can induce mitochondrial damage independently from the

transcriptional activity of p53. Since p53 also activates the

mitochondrial death pathway, the mitochondrion appears to

integrate the diverse pro-apoptotic mechanisms induced by

oncogenes. Studies in transgenic mice have revealed that

Bcl-2 (and/or Bcl-xl) over-expression and p53 mutations

50 Mol Cell Biochem (2011) 351:41–58

123

(or ARF loss) are selected independently during Myc-

induced lymphomagenesis [205–207].

The emerging knowledge about proteins that are

involved in apoptosis, including their 3-D structures and

biochemical mechanisms, has provided therapeutic ave-

nues by discovering molecules or targets which may

modulate apoptosis. A number of therapeutic approaches

are undergoing using an antisense RNA [208], various

small molecules [209–218, 222, 223], and peptidic com-

pounds [219–221, 224, 225] classified as potential thera-

peutics to target the pathway of apoptosis, as briefly

summarized in Table 2. Most of these therapeutics involve

the targeting of structurally defined multidomain of the

member of Bcl-2 family of protein. Some of these mole-

cules/agents are in clinical trials.

The combination of IAP antagonists with drugs that

target ErbB receptors promotes apoptosis thereby reduces

the cell turnover of breast cancer cells [226]. The apoptotic

response to most chemotherapeutic drugs in mammalian

cells involves the induction of mitochondrial pathway in

which mitochondrial membrane permeabilization con-

trolled by the Bcl-2 protein family, is induced. Other

strategies include the IAP proteins as therapeutic targets

that are expressed in the majority of human tumor through

the inhibition of cellular death and participation in sig-

naling pathways associated with malignancies [157, 227].

Further more, Table 2 has summarized the list of antisense

[228], small molecules [229–235], and peptidic compounds

[234, 236, 237] that are under investigation to develop

potential therapeutics against IAPs protein which are

involved in cancer.

Acknowledgments MSO and MN thank Medical Research Chair in

Ophthalmology funded by Dr. Nasser Al-Rasheed, College of Med-

icine, Kind Saud University for support. HA would like to thank Dr.

Nihal Ahmad, School of Medicine and Public Health, University of

Wiscosnsin, Madison for a cancer cell biology research fellowship.

The authors would also like to thank Ms. Crisalis Longanilla-Bautista

and Mr. Miaraj Siddiquei in helping with figures and proof reading.

References

1. Kerr JFR, Wyllie AH, Currie AR (1972) Apoptosis: a basic

biological phenomenon with wide-ranging implications in tissue

kinetics. Br J Cancer 26:239–257

2. Adams JM (2003) Ways of dying: multiple pathways to apop-

tosis. Genes Dev 17:2481–2495

3. Saikumar P, Dong Z, Mikhailov V, Denton M, Weinberg JM,

Venkatachalam MA (1999) Apoptosis: definition, mechanisms,

and relevance to disease. Am J Med 107:489–506

4. Gulbins E, Jekle A, Ferlinz K, Grassme H, Lang F (2000)

Physiology of apoptosis. Am J Physiol Renal Physiol 279:

605–615

5. Janssen O, Qian J, Linkermann A, Kabelitz D (2003) CD95

ligand—death factor and costimulatory molecule? Cell Death

Differ 10:1215–1225

6. Schutze S, Tchikov V, Schneider-Brachert W (2008) Regulation

of TNFR1 and CD95 signalling by receptor compartmentaliza-

tion. Nat Rev Mol Cell Biol 9:655–662

7. Ashkenazi A (2002) Targeting death and decoy receptors of the

tumour-necrosis factor superfamily. Nat Rev Cancer 2:420–430

8. Letai A (2006) Growth factor withdrawal and apoptosis: the

middle game. Mol Cell 21:728–730

Table 2 Compounds on high clinical development against the Bcl-2 and IAP’s family of proteins involved in the mitochondria-mediated

apoptosis pathway

Compound Class Mechanism (target) Reference

Targets for antiapoptotic Bcl-2 family of protein

Genasense (Bcl-2 antisense) Antisense Anti-apoptotic mRNA (Bcl-2) [208]

BH3I-1/BH3I-2; antimycin A3; compound 6;

apogossypol; theaflavanin; GX15-070; AT101: (-)

gossypol; polyphenol E; A-779024 (ABT 737);

HA14-1 analogs

Small molecules Anti-apoptotic inhibition (Bcl-2) [209–218]

CPM-1285 analogs; terphenyl derivative; SAHBs Peptide Anti-apoptotic inhibition (Bcl-2) [219–221]

Targets for pro-apoptotic Bcl-2 family of protein

3,6-dibromocarbazole piperazine derivatives of

2-propanol; 4-phenylsulfanyl-phenylamine derivatives

Small molecules Pro-apoptotic inhibition (BAX)/(BID) [222, 223]

Humanin peptides; Ku70 peptides Peptides Pro-apoptotic inhibition (BAX) [224, 225]

Targets for IAP’s family of protein

XIAP antisense (AEG35156/GEM640) Antisense Inhibition of XIAP [228]

Benzenesulfonamide derivatives; di/triphenylureas

(1396-11,12,34); embeline; compound 3/8/11; SM-

164; LBW242

Small molecules Inhibition of IAP [229–235]

Capped tripeptides containing unnatural amino acids;

Smac/penetratin fusion peptides

Peptide Inhibition of IAP [234, 236, 237]

Mol Cell Biochem (2011) 351:41–58 51

123

9. Zhang Y, Xing D, Liu L (2009) PUMA promotes Bax translo-

cation by both directly interacting with Bax and by competitive

binding to Bcl-X L during UV-induced apoptosis. Mol Biol Cell

20:3077–3087

10. Gentile M, Latonen L, Laiho M (2003) Cell cycle arrest and

apoptosis provoked by UV radiation-induced DNA damage are

transcriptionally highly divergent responses. Nucleic Acids Res

31:4779–4790

11. Stevenson MA, Pollock SS, Coleman CN, Calderwood SK

(1994) X-irradiation, phorbol esters, and H2O2 stimulate mito-

gen-activated protein kinase activity in NIH-3T3 cells through

the formation of reactive oxygen intermediates. Cancer Res

54:12–15

12. Dudeja V, Mujumdar N, Phillips P, Chugh R, Borja-Cacho D,

Dawra RK, Vickers SM, Saluja AK (2009) Heat shock pro-

tein 70 inhibits apoptosis in cancer cells through simultaneous

and independent mechanisms. Gastroenterology 136:1772–

1782

13. Solary E, Droin N, Bettaieb A, Corcos L, Dimanche-Boitrel MT,

Garrido C (2000) Positive and negative regulation of apoptotic

pathways by cytotoxic agents in hematologica. Leukemia 14:

1833–1849

14. Ahsan H, Reagan-Shaw S, Breur J, Ahmad N (2007) Sanguin-

arine induces apoptosis of human pancreatic carcinoma AsPC-1

and BxPC-3 cells via modulations in Bcl-2 family proteins.

Cancer Lett 249:198–208

15. Meng SJ, Yu LJ (2010) Oxidative stress, molecular inflamma-

tion and sarcopenia. Int J Mol Sci 11:1509–1526

16. Gu X, Song X, Dong Y, Cai H, Walters E, Zhang R, Pang X, Xie

T, Guo Y, Sridhar R, Califano JA (2008) Vitamin E succinate

induces ceramide-mediated apoptosis in head and neck squa-

mous cell carcinoma in vitro and in vivo. Clin Cancer Res

14:1840–1848

17. Lancellotti M, Pereira RF, Cury GG, Hollanda LM (2009)

Pathogenic and opportunistic respiratory bacteria-induced

apoptosis. Braz J Infect Dis 13:226–231

18. Sorensen CM (2004) Bcl-2 family members and disease. Bio-

chim Biophys Acta 1644:179–188

19. Thompson CB (1995) Apoptosis in the pathogenesis and treat-

ment of disease. Science 267:1456–1462

20. Hanahan D, Weinberg RA (2000) The hallmarks of cancer. Cell

100:57–70

21. Bidere N, Su HC, Lenardo MJ (2006) Genetic disorders of

programmed cell death in the immune system. Annu Rev

Immunol 24:321–352

22. Mersich S, Gadaleta P (2003) Nuevas estrategias terapeuticas

basadas en apoptosis y virus. Acta Bioquımica Clınica Latino-

americana 37:13–21

23. Fisher DE (1994) Apoptosis in cancer therapy: crossing the

threshold. Cell 78:539–542

24. Johnstone RW, Ruefli AA, Lowe SW (2002) Apoptosis:

a link between cancer genetics and chemotherapy. Cell 108:

153–164

25. Pecina-Slaus N (2009) Genetic and molecular insights into

apoptosis. Acta Med Croatica 63(Suppl 2):13–19

26. Schaffitzel E, Hertweck M (2006) Recent aging research in

Caenorhabditis elegans. Exp Gerontol 41:557–563

27. Danial NN, Korsmeyer SJ (2004) Cell death: critical control

points. Cell 116(2):205–219

28. Horvitz HR (1999) Genetic control of programmed cell death in

the nematode Caenorhabditis elegans. Cancer Res 59:1701S–

1706S

29. McDonnell TJ, Deane N, Platt FM, Nunez G, Jaeger U,

McKearn JP, Korsmeyer SJ (1989) Bcl-2-immunoglobulin

transgenic mice demonstrate extended B cell survival and fol-

licular lymphoproliferation. Cell 57:79–88

30. Vaux DL, Cory S, Adams JM (1988) Bcl-2 gene promotes

haemopoietic cell survival and cooperates with c-myc to

immortalize pre-B cells. Nature 335:440–442

31. Vaux DL, Weissman IL, Kim SK (1992) Prevention of pro-

grammed cell death in Caenorhabditis elegans by human bcl-2.

Science 258:1955–1957

32. Hengartner MO, Horvitz HR (1994) C. elegans cell survival

gene ced-9 encodes a functional homolog of the mammalian

proto-oncogene bcl-2. Cell 76:665–676

33. Mohamad N, Gutierrez A, Nunez M, Cocca C, Martın G, Cricco

G, Medina V, Rivera E, Bergoc R (2005) Mitochondrial apop-

totic pathways. Biocell 29:149–161

34. Fan TJ, Han LH, Cong RS, Liang J (2005) Caspase family

proteases and apoptosis. Acta Biochimica et Biophysica Sinica

37:719–727

35. Li J, Yuan J (2008) Caspases in apoptosis and beyond. Onco-

gene 27(48):6194–6206

36. Petros AM, Olejniczak ET, Fesik SW (2004) Structural biology

of the Bcl-2 family of proteins. Biochim Biophys Acta

1644:83–94

37. Germain M, Shore GC (2003) Cellular distribution of Bcl-2

family proteins. Sci STKE 173:pe10

38. Budd R (2001) Activation-induced cell death. Curr Opin

Immunol 13:356–362

39. Wang K, Yin XM, Chao DT, Milliman CL, Korsmeyer SJ

(1996) BID: a novel BH3 domain-only death agonist. Genes

Dev 10:2859–2869

40. Brunelle JK, Letai A (2009) Control of mitochondrial apoptosis

by the Bcl-2 family. J Cell Sci 122:437–441

41. Puthalakath H, Strasser A (2002) Keeping fillers on a tight leash:

transcriptional and post-translational control of the proapoptotic

activity of BH3-only proteins. Cell Death Differ 9:505–512

42. Ogilvy S, Metcalf D, Print CG, Bath ML, Harris AW, Adams

JM (1999) Constitutive Bcl-2 expression throughout the hema-

topoietic compartment affects multiple lineages and enhances

progenitor cell survival. Proc Natl Acad Sci USA 96:

14943–14948

43. Veis DJ, Sorenson CM, Shutter JR, Korsmeyer SJ (1993) Bcl-2-

deficient mice demonstrate fulminant lymphoid apoptosis,

polycystic kidneys, and hypopigmented hair. Cell 75:229–240

44. Ross AJ, Waymire KG, Moss JE, Parlow AF, Skinner MK,

Russell LD, MacGregor GR (1998) Testicular degeneration in

Bcl-w-deficient mice. Nat Genet 18:251–256

45. Hamasaki A, Sendo F, Nakayama K, Ishida N, Negishi I,

Nakayama K, Hatakeyama S (1998) Accelerated neutrophil

apoptosis in mice lacking A1-a, a subtype of the Bcl-2-related

A1 gene. J Exp Med 188:1985–1992

46. Rinkenberger JL, Horning S, Klocke B, Roth K, Korsmeyer SJ

(2000) Mcl-1 deficiency results in peri-implantation embryonic

lethality. Genes Dev 14:23–27

47. Motoyama N, Wang F, Roth KA, Sawa H, Nakayama K,

Nakayama K, Negishi I, Senju S, Zhang Q, Fujii S (1995)

Massive cell death of immature hematopoietic cells and neurons

in Bcl-x deficient mice. Science 267:1506–1510

48. Youle RJ, Strasser A (2008) The BCL-2 protein family:

opposing activities that mediate cell death. Nat Rev Mol Cell

Biol 9:47–59

49. Brooks C, Wei Q, Feng L, Dong G, Tao Y, Mei L, Xie ZJ, Dong

Z (2007) Bak regulates mitochondrial morphology and pathol-

ogy during apoptosis by interacting with mitofusins. Proc Natl

Acad Sci USA 104:11649–11654

50. Mikhailov V et al (2003) Association of Bax and Bak homoo-

ligomers in mitochondria. Bax requirement for Bak reorgani-

zation and cytochrome c release. J Biol Chem 278:5367–5376

51. Basanez G, Sharpe JC, Galanis J, Brandt TB, Hardwick JM,

Zimmerberg J (2002) Bax-type apoptotic proteins porate pure

52 Mol Cell Biochem (2011) 351:41–58

123

lipid bilayers through a mechanism sensitive to intrinsic

monolayer curvature. J Biol Chem 277:49360–49365

52. Kuwana T, Mackey MR, Perkins G, Ellisman MH, Latterich M,

Schneiter R, Green DR, Newmeyer DD (2002) Bid, Bax, and

lipids cooperate to form supramolecular openings in the outer

mitochondrial membrane. Cell 111:331–342

53. Yethon JA, Epand RF, Leber B, Epand RM, Andrews DW

(2003) Interaction with a membrane surface triggers a reversible

conformational change in Bax normally associated with induc-

tion of apoptosis. J Biol Chem 278:48935–48941

54. Wei MC, Zong WX, Cheng EH, Lindsten T, Panoutsakopoulou

V, Ross AJ, Roth KA, MacGregor GR, Thompson CB, Kors-

meyer SJ (2001) Proapoptotic BAX and BAK: a requisite

gateway to mitochondrial dysfunction and death. Science

292:727–730

55. Wang X (2001) The expanding role of mitochondria in apop-

tosis. Genes Dev 15:922–2933

56. Lovell JF, Billen LP, Bindner S, Shamas-Din A, Fradin C, Leber

B, Andrews DW (2008) Membrane binding by tBid initiates an

ordered series of events culminating in membrane permeabili-

zation by Bax. Cell 135(6):1074–1084

57. Letai A, Bassik MC, Walensky LD, Sorcinelli MD, Weiler S,

Korsmeyer SJ (2002) Distinct BH3 domains either sensitize or

activate mitochondrial apoptosis, serving as prototype cancer

therapeutics. Cancer Cell 2:183–192

58. Chipuk JE, Kuwana T, Bouchier-Hayes L, Droin NM, New-

meyer DD, Schuler M, Green DR (2004) Direct activation of

Bax by p53 mediates mitochondrial membrane permeabilization

and apoptosis. Science 303:1010–1014

59. Kim H, Rafiuddin-Shah M, Tu HC, Jeffers JR, Zambetti GP,

Hsieh JJ, Cheng EH (2006) Hierarchical regulation of mito-

chondrion-dependent apoptosis by BCL-2 subfamilies. Nat Cell

Biol 8:1348–1358

60. Pagliari LJ, Kuwana T, Bonzon C, Newmeyer DD, Tu S, Beere

HM, Green DR (2005) The multidomain proapoptotic molecules

Bax and Bak are directly activated by heat. Proc Natl Acad Sci

USA 102:17975–17980

61. Gavathiotis E, Suzuki M, Davis ML, Pitter K, Bird GH, Katz

SG, Tu HC, Kim H, Cheng EH, Tjandra N, Walensky LD (2008)

BAX activation is initiated at a novel interaction site. Nature

455:1076–1081

62. Willis SN, Fletcher JI, Kaufmann T, van Delft MF, Chen L,

Czabotar PE, Ierino H, Lee EF, Fairlie WD, Bouillet P, Strasser

A, Kluck RM, Adams JM, Huang DC (2007) Apoptosis initiated

when BH3 ligands engage multiple Bcl-2 homologs, not Bax or

Bak. Science 315:856–859

63. Fletcher JI, Meusburger S, Hawkins CJ, Riglar DT, Lee EF,

Fairlie WD, Huang DC, Adams JM (2008) Apoptosis is trig-

gered when prosurvival Bcl-2 proteins cannot restrain Bax. Proc

Natl Acad Sci USA 105:18081–18087

64. Cheng EH, Sheiko TV, Fisher JK, Craigen WJ, Korsmeyer SJ

(2003) VDAC2 inhibits BAK activation and mitochondrial

apoptosis. Science 301:513–517

65. Li H, Zhu H, Xu CJ, Yuan J (1998) Cleavage of BID by caspase

8 mediates the mitochondrial damage in the Fas pathway of

apoptosis. Cell 94:491–501

66. Luo X, Budihardjo I, Zou H, Slaughter C, Wang X (1998) Bid, a

Bcl2 interacting protein, mediates cytochrome c release from

mitochondria in response to activation of cell surface death

receptors. Cell 94:481–490

67. Billen LP, Shamas-Din A, Andrews DW (2008) Bid: a Bax-like

BH3 protein. Oncogene 27(Suppl 1):S93–S104

68. Oh KJ, Barbuto S, Pitter K, Morash J, Walensky LD, Korsmeyer

SJ (2006) A membrane-targeted BID BCL-2 homology 3 pep-

tide is sufficient for high potency activation of BAX in vitro.

J Biol Chem 281:36999–37008

69. Eskes R, Desagher S, Antonsson B, Martinou JC (2000) Bid

induces the oligomerization and insertion of Bax into the outer

mitochondrial membrane. Mol Cell Biol 20:929–935

70. Balakrishnan G, Hu Y, Oyerinde OF, Su J, Groves JT, Spiro TG

(2007) A conformational switch to b-sheet structure in cyto-

chrome c leads to heme exposure. Implications for cardiolipin

peroxidation and apoptosis. J Am Chem Soc 129:504–505

71. Kim TH et al (2004) Bid–cardiolipin interaction at mitochon-

drial contact site contributes to mitochondrial cristae reorgani-

zation and cytochrome c release. Mol Biol Cell 15:3061–3072

72. Giordano A et al (2005) tBid induces alterations of mitochon-

drial fatty acid oxidation flux by malonyl-CoA-independent

inhibition of carnitine palmitoyltransferase-1. Cell Death Differ

12:603–613

73. Tyurin VA et al (2007) Interactions of cardiolipin andlyso-car-

diolipins with cytochrome c and tBid: conflict or assistance in

apoptosis. Cell Death Differ 14:872–875

74. Orrenius S, Gogvadze V, Zhivotovsky B (2007) Mitochondrial

oxidative stress: implications for cell death. Annu Rev Phar-

macol Toxicol 47:143–183

75. Zha J, Harada H, Yang E, Jockel J, Korsmeyer SJ (1996) Serine

phosphorylation of death agonist BAD in response to survival

factor results in binding to 14-3-3 not BCL-X(L). Cell 87:

619–628

76. Lodish H, Berk A, Zipursky SL, Matsudaira P, Baltimore D,

Darnell JE, Freeman WH (2000) Molecular cell biology, 4th

edn. W. H. Freeman & Co, New York, Chapter 23–28

77. Datta SR, Dudek H, Tao X, Masters S, Fu H, Gotoh Y,

Greenberg ME (1997) Akt phosphorylation of BAD couples

survival signals to the cell-intrinsic death machinery. Cell 91:

231–241

78. Tommasini I, Cerioni L, Palomba L, Cantoni O (2008) Prosta-

glandin E2 signals monocyte/macrophage survival to perox-

ynitrite via protein kinase A converging in bad phosphorylation

with the protein kinase C alpha-dependent pathway driven by

5-hydroxyeicosatetraenoic acid. J Immunol 181:5637–5645

79. Grund K, Ahmadi R, Jung F, Funke V, Gdynia G, Benner A,

Sykora J, Walczak H, Joos S, Felsberg J, Reifenberger G,

Wiestler OD, Herold-Mende C, Roth W (2008) Troglitazone-

mediated sensitization to TRAIL-induced apoptosis is regulated

by proteasome-dependent degradation of FLIP and ERK1/2-

dependent phosphorylation of BAD. Cancer Biol Ther

7:1982–1990

80. Ahn S, Kim J, Hara MR, Ren XR, Lefkowitz RJ (2009) {Beta}-

arrestin-2 mediates anti-apoptotic signaling through regulation

of BAD phosphorylation. J Biol Chem 284:8855–8865

81. Puthalakath H, Villunger A, O’Reilly LA, Beaumont JG,

Coultas L, Cheney RE, Huang DC, Strasser A (2001) Bmf: a

pro-apoptotic BH3-only protein regulated by interaction with the

myosin V actin motor complex, activated by anoikis. Science

293:1829–1832

82. Lei K, Davis RJ (2003) JNK phosphorylation of Bim-related

members of the Bcl2 family induces Bax-dependent apoptosis.

Proc Natl Acad Sci USA 100:2432–2437

83. Okuno S, Saito A, Hayashi T, Chan PH (2004) The c-Jun

N-terminal protein kinase signaling pathway mediates Bax

activation and subsequent neuronal apoptosis through interac-

tion with Bim after transient focal cerebral ischemia. J Neurosci

24:7879–7887

84. Bouillet P, Metcalf D, Huang DC, Tarlinton DM, Kay TW,

Kontgen F, Adams JM, Strasser A (1999) Proapoptotic Bcl-2

relative Bim required for certain apoptotic responses, leukocyte

homeostasis, and to preclude autoimmunity. Science 286:

1735–1738

85. Bouillet P, Purton JF, Godfrey DI, Zhang LC, Coultas L,

Puthalakath H, Pellegrini M, Cory S, Adams JM, Strasser A

Mol Cell Biochem (2011) 351:41–58 53

123

(2002) BH3-only Bcl-2 family member Bim is required for

apoptosis of autoreactive thymocytes. Nature 415:922–926

86. Putcha GV, Moulder KL, Golden JP, Bouillet P, Adams JA,

Strasser A, Johnson EM (2001) Induction of Bim, a proapoptotic

BH3- only Bcl-2 family member, is critical for neuronal apop-

tosis. Neuron 29:615–628

87. Zhang L, Xing D, Chen M (2008) Bim(L) displacing Bcl-

x(L) promotes Bax translocation during TNFalpha-induced

apoptosis. Apoptosis 13:950–958

88. Wang X, Xing D, Liu L, Chen WR (2009) BimL directly neu-

tralizes Bcl-xL to promote Bax activation during UV-induced

apoptosis. FEBS Lett 583:1873–1879

89. Zhang Y, Adachi M, Kawamura R, Zou HC, Imai K, Hareyama

M, Shinomura Y (2006) Bmf contributes to histone deacetylase

inhibitor-mediated enhancing effects on apoptosis after ionizing

radiation. Apoptosis 11:1349–1357

90. Zhang Y, Adachi M, Kawamura R, Imai K (2006) Bmf is a

possible mediator in histone deacetylase inhibitors FK228 and

CBHA-induced apoptosis. Cell Death Differ 13:129–140

91. Ramjaun AR, Tomlinson S, Eddaoudi A, Downward J (2007)

Upregulation of two BH3-only proteins, Bmf and Bim, during

TGF beta-induced apoptosis. Oncogene 26:970–981

92. Yakovlev AG, Giovanni SD, Wang G et al (2004) Bok and Noxa

are essential mediators of p53-dependent apoptosis. J Biol Chem

279:28367–28374

93. Oda E, Ohki R, Murasawa H et al (2000) Noxa, a BH3- only

member of the Bcl-2 family and candidate mediator of p53-

induced apoptosis. Science 288:1053–1058

94. Villunger A, Michalak EM, Coultas L et al (2003) p53- and

drug-induced apoptotic responses mediated by BH3-only pro-

teins Puma and Noxa. Science 302:1036–1038

95. Karst AM, Li G (2007) BH3-only proteins in tumorigenesis and

malignant melanoma. Cell Mol Life Sci 64:318–330

96. Ming L, Wang P, Bank A et al (2006) Puma dissociates Bax and

Bcl-XL to induce apoptosis in colon cancer cells. J Biol Chem

281:16034–16042

97. Wyttenbach A, Tolkovsky AM (2006) The BH3-only protein

Puma is both necessary and sufficient for neuronal apoptosis

induced by DNA damage in sympathetic neurons. J Neurochem

96:1213–1226

98. Jabbour AM, Heraud JE, Daunt CP, Kaufmann T, Sandow J,

O’Reilly LA, Callus BA, Lopez A, Strasser A, Vaux DL, Ekert

PG (2009) Puma indirectly activates Bax to cause apoptosis in

the absence of Bid or Bim. Cell Death Differ 16:555–563

99. Liu Z, Lu H, Shi H et al (2005) Puma overexpression induces

reactive oxygen species generation and proteasome- mediated

stathmin degradation in colorectal cancer cells. Cancer Res

65:1647–1654

100. Hemann MT, Zilfou JT, Zhao Z et al (2004) Suppression of

tumorigenesis by the p53 target Puma. Proc Natl Acad Sci USA

101:9333–9338

101. Adams JM, Cory S (2007) The Bcl-2 apoptotic switch in cancer

development and therapy. Oncogene 26:1324–1337

102. Riedl SJ, Salvesen GS (2007) The apoptosome: signalling

platform of cell death. Nat Rev Mol Cell Biol 8:405–413

103. Strasser A, O’Connor L, Dixit VM (2000) Apoptosis signaling.

Annu Rev Biochem 69:217–245

104. Strasser A, Harris AW, Huang DCS, Krammer PH, Cory S

(1995) Bcl-2 and Fas/APO-1 regulate distinct pathways to

lymphocyte apoptosis. EMBO J 14:6136–6147

105. Li P et al (1997) Cytochrome c and dATP-dependent formation

of Apaf-1/caspase-9 complex initiates an apoptotic protease

cascade. Cell 91:479–489

106. Ghavami S, Hashemi M, Ande SR, Yeganeh B, Xiao W,

Eshraghi M, Bus CJ, Kadkhoda K, Wiechec E, Halayko AJ, Los

M (2009) Apoptosis and cancer: mutations within caspase

genes. J Med Genet 46:497–510

107. Locksley RM, Killeen N, Lenardo MJ (2001) The TNF and TNF

receptor superfamilies: integrating mammalian biology. Cell

104:487–501

108. Wajant H (2002) The Fas signaling pathway: more than a par-

adigm. Science 296:1635–1636

109. Chen G, Goeddel D (2002) TNF-1 signaling: a beautiful path-

way. Science 296:1634–1635

110. Yang JK, Wang L, Zheng L, Wan F, Ahmed M, Lenardo MJ,

Wu H (2005) Crystal structure of MC159 reveals molecular

mechanism of DISC assembly and FLIP inhibition. Mol Cell

20:939–949

111. Korsmeyer SJ, Wei MC, Saito M, Weiler S, Oh KJ, Schlesinger

PH (2000) Pro-apoptotic cascade activates BID, which oligo-

merizes BAK or BAX into pores that result in the release of

cytochrome c. Cell Death Differ 7:1166–1173

112. Lartigue L, Kushnareva Y, Seong Y, Lin H, Faustin B,

Newmeyer DD (2009) Caspase-independent mitochondrial cell

death results from loss of respiration, not cytotoxic protein

release. Mol Biol Cell 20:4871–4884

113. Pellegrini L, Scorrano L (2007) A cut short to death: Parl and

Opa1 in the regulation of mitochondrial morphology and

apoptosis. Cell Death Differ 14:1275–1284

114. Liu X, Kim CN, Yang J, Jemmerson R, Wang X (1996)

Induction of apoptotic program in cell-free extracts: requirement

for dATP and cytochrome c. Cell 86:147–157

115. Li F et al (1997) Cell-specific induction of apoptosis by

microinjection of cytochrome c. Bcl-xL has activity independent

of cytochrome c release. J Biol Chem 272:30299–30305

116. Zhivotovsky B, Orrenius S, Brustugun OT, Doskeland SO

(1998) Injected cytochrome c induces apoptosis. Nature

391:449–450

117. Newmeyer DD, Farschon DM, Reed JC (1994) Cell-free apop-

tosis in Xenopus egg extracts: inhibition by Bcl-2 and require-

ment for an organelle fraction enriched in mitochondria. Cell

79:353–364

118. Kluck RM, Bossy-Wetzel E, Green DR, New meyer DD (1997)

The release of cytochrome c from mitochondria: a primary site

for Bcl-2 regulation of apoptosis. Science 275:1132–1136

119. Glazunova VA, Shtil AA (2008) Mitochondrial mechanisms of

apoptosis in response to DNA damage. Mol Biol (Mosk)

42:765–771

120. Ravi D, Das KC (2004) Redox-cycling of anthracyclines by

thioredoxin system: increased superoxide generation and DNA

damage. Cancer Chemother Pharmacol 54:449–458

121. Yang J, Liu X, Bhalla K, Kim CN, Ibrado AM, Cai J, Peng TI,

Jones DP, Wang X (1997) Prevention of apoptosis by Bcl-2:

release of cytochrome c from mitochondria blocked. Science

275:1129–1132

122. Bajt ML, Cover C, Lemasters JJ, Jaeschke H (2006) Nuclear

translocation of endonuclease G and apoptosis-inducing factor

during acetaminophen-induced liver cell injury. Toxicol Sci

94:217–225

123. Li LY, Luo X, Wang X (2001) Endonuclease G (EndoG) is an

apoptotic DNAse when released from mitochondria. Nature

412:95–99

124. Zhang J, Liu X, Scherer DC, van Kaer L, Wang X, Xu M (1998)

Resistance to DNA fragmentation and chromatin condensation

in mice lacking the DNA fragmentation factor. Proc Natl Acad

Sci 95:12480–12485

125. Ohsato T, Ishihara N, Muta T, Umeda S, Ikeda S, Mihara K,

Hamasaki N, Kang D (2002) Mammalian mitochondrial endo-

nuclease G digestion of R-loops and localization in intermem-

brane space. Eur J Biochem 269:5765–5770

54 Mol Cell Biochem (2011) 351:41–58

123

126. van Loo G, Schotte P, van Gurp M, Demol H, Hoorelbeke B,

Gevaert K, Rodriguez I, Ruiz-Carrillo A, Vandekerckhove J,

Declercq W, Beyaert R, Vandenabeele P (2001) Endonuclease

G: a mitochondrial protein released in apoptosis and involved in

caspase-independent DNA degradation. Cell Death Differ

8:1136–1142

127. Widlak P, Li LY, Wang X, Garrard WT (2001) Action of

recombinant human apoptotic endonuclease G on naked DNA

and chromatin substrates: cooperation with exonuclease and

DNase. I. J Biol Chem 276:48404–48409

128. Susin SA, Lorenzo HK, Zamzami N, Marzo I, Snow BE,

Brothers GM, Mangion J, Jacotot E, Costantini P, Loeffler M,

Larochette N, Goodlett DR, Aebersold R, Siderovski DP,

Penninger JM, Kroemer G (1999) Molecular characterization of

mitochondrial apoptosis-inducing factor. Nature 397:441–446

129. Kondo K, Obitsu S, Ohta S, Matsunami K, Otsuka H, Teshima R

(2010) Poly(ADP-ribose) polymerase (PARP)-1-independent

apoptosis-inducing factor (AIF) release and cell death are

induced by eleostearic acid and blocked by a-tocopherol and

MEK inhibition. J Biol Chem 285:13079–13091

130. Miramar MD, Costantini P, Ravagnan L, Saraiva LM, Haouzi D,

Brothers G, Penninger JM, Peleato ML, Kroemer G, Susin SA

(2001) NADH oxidase activity of mitochondrial apoptosis-

inducing factor. J Biol Chem 276:16391–16398

131. Joza N, Pospisilik JA, Hangen E, Hanada T, Modjtahedi N,

Penninger JM, Kroemer G (2009) AIF: not just an apoptosis-

inducing factor. Ann N Y Acad Sci 1171:2–11

132. Joza N, Susin SA, Daugas E, Stanford WL, Cho SK, Li CY,

Sasaki T, Elia AJ, Cheng HY, Ravagnan L, Ferri KF, Zamzami

N, Wakeham A, Hakem R, Yoshida H, Kong YY, Mak TW,

Zuniga-Pflucker JC, Kroemer G, Penninger JM (2001) Essential

role of the mitochondrial apoptosis-inducing factor in pro-

grammed cell death. Nature 410:549–554

133. Schulthess FT, Katz S, Ardestani A, Kawahira H, Georgia S,

Bosco D, Bhushan A, Maedler K (2009) Deletion of the mito-

chondrial flavoprotein apoptosis inducing factor (AIF) induces

b-cell apoptosis and impairs b-cell mass. PLoS One 4:4394

134. Cande C, Cecconi F, Dessen P, Kroemer G (2002) Apoptosis-

inducing factor (AIF): key to the conserved caspase-independent

pathways of cell death? J Cell Sci 115:4727–4734

135. Yu W, Gubkina O, Mechawar N, Elwell D, Quirion R, Krantic S

(2009) Expression of apoptosis-inducing factor (AIF) in the

aged rat brain. Neurobiol Aging 32(1):179–180

136. Susin SA, Daugas E, Ravagnan L, Samejima K, Zamzami N,

Loeffler M, Costantini P, Ferri KF, Irinopoulou T, Prevost MC,

Brothers G, Mak TW, Penninger J, Earnshaw WC, Kroemer G

(2000) Two distinct pathways leading to nuclear apoptosis.

J Exp Med 192:571–580

137. Pallast S, Arai K, Pekcec A, Yigitkanli K, Yu Z, Wang X, Lo

EH, Leyen KV (2010) Increased nuclear apoptosis-inducing

factor after transient focal ischemia: a 12/15-lipoxygenase-

dependent organelle damage pathway. J Cereb Blood Flow

Metab 30:1157–1167

138. Verhagen AM, Ekert PG, Pakusch M, Silke J, Connolly LM,

Reid GE, Moritz RL, Simpson RJ, Vaux DL (2000) Identifica-

tion of DIABLO, a mammalian protein that promotes apoptosis

by binding to and antagonizing IAP proteins. Cell 102:43–53

139. Kominsky DJ, Bickel RJ, Tyler KL (2002) Reovirus-induced

apoptosis requires mitochondrial release of Smac/DIABLO and

involves reduction of cellular inhibitor of apoptosis protein

levels. J Virol 76:11414–11424

140. Wilkinson JC, Wilkinson AS, Scott FL, Csomos RA, Salvesen

GS, Duckett CS (2004) Neutralization of Smac/Diablo by

inhibitors of apoptosis (IAPs). A caspase-independent mecha-

nism for apoptotic inhibition. J Biol Chem 279:51082–51090

141. Martinez-Ruiz G, Maldonado V, Ceballos-Cancino G, Grajeda

JP, Melendez-Zajgla J (2008) Role of Smac/DIABLO in cancer

progression. J Exp Clin Cancer Res 26:48

142. Arellano-Llamas A, Garcia FJ, Perez D, Cantu D, Espinosa M,

De la Garza JG, Maldonado V, Melendez-Zajgla J (2006) High

Smac/DIABLO expression is associated with early local recur-

rence of cervical cancer. BMC Cancer 6:256

143. Kohli M, Yu J, Seaman C, Bardelli A, Kinzler KW, Vogelstein

B, Lengauer C, Zhang L (2004) SMAC/Diablo-dependent

apoptosis induced by nonsteroidal antiinflammatory drugs

(NSAIDs) in colon cancer cells. Proc Natl Acad Sci USA

101:16897–16902

144. Augello C, Caruso L, Maggioni M, Donadon M, Montorsi M,

Santambrogio R, Torzilli G, Vaira V, Pellegrini C, Roncalli M,

Coggi G, Bosari S (2009) Inhibitors of apoptosis proteins (IAPs)

expression and their prognostic significance in hepatocellular

carcinoma. BMC Cancer 9:125

145. Gray CW, Ward RV, Karran E, Turconi S, Rowles A,

Viglienghi D, Southan C, Barton A, Fantom KG, West A,

Savopoulos J, Hassan NJ, Clinkenbeard H, Hanning C, Ame-

gadzie B, Davis JB, Dingwall C, Livi GP, Creasy CL (2000)

Characterization of human HtrA2, a novel serine protease

involved in the mammalian cellular stress response. Eur J Bio-

chem 267:5699–5710

146. Suzuki Y, Imai Y, Nakayama H, Takahashi K, Takio K,

Takahashi R (2001) A serine protease, HtrA2, is released from

the mitochondria and interacts with XIAP, inducing cell death.

Mol Cell 8:613–621

147. Verhagen AM, Silke J, Ekert PG, Pakusch M, Kaufmann H,

Connolly LM, Day CL, Tikoo A, Burke R, Wrobel C, Moritz

RL, Simpson RJ, Vaux DL (2002) HtrA2 promotes cell death

through its serine protease activity and its ability to antagonize

inhibitor of apoptosis proteins. J Biol Chem 277:445–454

148. Martins LM, Iaccarino I, Tenev T, Gschmeissner S, Totty NF,

Lemoine NR, Savopoulos J, Gray CW, Creasy CL, Dingwall C,

Downward J (2002) The serine protease Omi/HtrA2 regulates

apoptosis by binding XIAP through a repear-like motif. J Biol

Chem 277:439–444

149. Martins LM, Morrison A, Klupsch K, Fedele V, Moisoi N,

Teismann P, Abuin A, Grau E, Geppert M, Livi GP, Creasy CL,

Martin A, Hargreaves I, Heales SJ, Okada H, Brandner S, Schulz

JB, Mak T, Downward J (2004) Neuroprotective role of the

reaper-related serine protease HtrA2/Omi revealed by targeted

deletion in mice. Mol Cell Biol 24:9848–9862

150. Balakrishnan MP, Cilenti L, Mashak Z, Popat P, Alnemri ES,

Zervos AS (2009) THAP5 is a human cardiac-specific inhibitor

of cell cycle that is cleaved by the proapoptotic Omi/HtrA2

protease during cell death. Am J Physiol Heart Circ Physiol

297:H643–H653

151. Pruefer FG, Lizarraga F, Maldonado V, Melendez-Zajgla J

(2008) Participation of Omi Htra2 serine-protease activity in the

apoptosis induced by cisplatin on SW480 colon cancer cells.

J Chemother 20:348–354

152. Ding X, Patel M, Shen D, Herzlich AA, Cao X, Villasmil R,

Klupsch K, Tuo J, Downward J, Chan CC (2009) Enhanced

HtrA2/Omi expression in oxidative injury to retinal pigment

epithelial cells and murine models of neurodegeneration. Invest

Ophthalmol Vis Sci 50:4957–4966

153. Trencia A, Fiory F, Maitan MA, Vito P, Barbagallo AP, Perfetti

A, Miele C, Ungaro P, Oriente F, Cilenti L, Zervos AS,

Formisano P, Beguinot F (2004) Omi/HtrA2 promotes cell death

by binding and degrading the anti-apoptotic protein ped/pea-15.

J Biol Chem 279:46566–46572

154. Hu XY, Chen XC, Zhu ZH, Chen CH, Zeng FQ, Lu GC (2006)Effects of Omi/HtrA2 on expression of anti-apoptotic protein

Mol Cell Biochem (2011) 351:41–58 55

123

PED/PEA-15 and apoptosis of prostate cancer cell line PC-3. Ai

Zheng 25:677–682

155. Krick S, Shi S, Ju W, Faul C, Tsai SY, Mundel P, Bottinger EP

(2008) Mpv17l protects against mitochondrial oxidative stress

and apoptosis by activation of Omi/HtrA2 protease. Proc Natl

Aca. Sci USA 105:14106–14111

156. Salvesen GS, Duckett CS (2002) IAP proteins: blocking the road

to death’s door. Nat Rev Mol Cell Biol 3:401–410

157. Vucic D, Fairbrother WJ (2007) The inhibitor of apoptosis

proteins as therapeutic targets in cancer. Clin Cancer Res

13:5995–6000

158. Srinivasula SM, Hegde R, Saleh A, Datta P, Shiozaki E, Chai J,

Lee RA, Robbins PD, Fernandes-Alnemri T, Shi Y, Alnemri ES

(2001) A conserved XIAP-interaction motif in caspase-9 and

Smac/DIABLO regulates caspase activity and apoptosis. Nature

410:112–116

159. Srinivasula SM, Gupta S, Datta P, Zhang Z, Hegde R, Cheong

N, Fernandes-Alnemri T, Alnemri ES (2003) Inhibitor of

apoptosis proteins are substrates for the mitochondrial serine

protease Omi/HtrA2. J Biol Chem 278:31469–31472

160. Huang Y, Park YC, Rich RL, Segal D, Myszka DG, Wu H

(2001) Structural basis of caspase inhibition by XIAP: differ-

ential roles of the linker versus the BIR domain. Cell 104:81–90

161. Eckelman BP, Salvesen GS (2006) The human antiapoptotic

proteins cIAP1 and cIAP2 bind but do not inhibit caspases.

J Biol Chem 281:3254–3260

162. Zobel K, Wang L, Varfolomeev E, Franklin MC, Elliott LO,

Wallweber HJ, Okawa DC, Flygare JA, Vucic D, Fairbrother

WJ, Deshayes K (2006) Design, synthesis, and biological

activity of apotent Smac mimetic that sensitizes cancer cells to

apoptosis by antagonizing IAPs. ACSChemBiol 1:525–533

163. Kempkensteffen C, Hinz S, Christoph F, Krause H, Magheli A,

Schrader M, Schostak M, Miller K, Weikert S (2008) Expression

levels of the mitochondrial IAP antagonists Smac/DIABLO and

Omi/HtrA2 in clear-cell renal cell carcinomas and their prog-

nostic value. J Cancer Res Clin Oncol 134:543–550

164. Wagener N, Crnkovic-Mertens I, Vetter C, Macher-Goppinger

S, Bedke J, Grone EF, Zentgraf H, Pritsch M, Hoppe-Seyler K,

Buse S, Haferkamp A, Autschbach F, Hohenfellner M, Hoppe-

Seyler F (2007) Expression of inhibitor of apoptosis protein

Livin in renal cell carcinoma and non-tumorous adult kidney. Br

J Cancer 97:1271–1276

165. Fulda S (2008) Targeting inhibitor of apoptosis proteins (IAPs)

for cancer therapy. Anticancer Agents Med Chem 8:533–539

166. Wang J, Lenardo MJ (2000) Roles of caspases in apoptosis,

development, and cytokine maturation revealed by homozygous

gene deficiencies. J Cell Sci 113(Pt 5):753–757

167. Shi Y (2002) Mechanisms of caspase activation and inhibition

during apoptosis. Mol. Cell 9:459–470

168. Los M, van de Craen M, Penning CL, Schenk H, Westendorp M,

Baeuerle PA, Droge W, Krammer PH, Fiers W, Schulze-Osthoff

K (1995) Requirement of an ICE/Ced-3 protease for Fas/Apo-

1–1mediated apoptosis. Nature 375:81–83

169. Thornberry NA, Lazebnik Y (1998) Caspases: enemies within.

Science 281:1312–1316

170. Sprick MR and Walczak H (2004) The interplay between the

Bcl-2 family and death receptor-mediated apoptosis. Biochim

Biophys Acta 1644:125–132

171. Bonzon C, Bouchier-Hayes L, Pagliari LJ, Green DR,

Newmeyer DD (2006) Caspase-2-induced apoptosis requires bid

cleavage: a physiological role for bid in heat shock-induced

death. Mol Biol Cell 17:2150–2157

172. Shi Y (2008) Apoptosome assembly Methods. Enzymol

442:141–156

173. Bao Q, Shi Y (2007) Apoptosome: a platform for the activation

of initiator caspases. Cell Death Differ 14:56–65

174. Ow YP, Green DR, Hao Z, Mak TW (2008) Cytochrome c:

functions beyond respiration. Nat Rev Mol Cell Biol 9:532–542

175. Peter ME, Krammer PH (2003) The CD95(APO-1/Fas) DISC

and beyond. Cell Death Differ 10:26–35

176. Kischkel FC, Hellbardt S, Behrmann I, Germer M, Pawlita M,

Krammer PH, peter ME (1995) Cytotoxicity-dependent APO-1

(Fas/CD95)-associated proteins form a death-inducing signaling

complex (DISC) with the receptor. EMBO J 14:5579–5588

177. Yan N, Shi Y (2005) Mechanisms of apoptosis through struc-

tural biology. Annu Rev Cell Dev Biol 21:35–56

178. Donepudi M, Mac Sweeney A, Briand C, Grutter MG (2003)

Insights into the regulatory mechanism for caspase-8 activation.

Mol Cell 11:543–549

179. Davis AR, Lotocki G, Marcillo AE, Dietrich WD, Keane RW

(2007) FasL, Fas, and death-inducing signaling complex (DISC)

proteins are recruited to membrane rafts after spinal cord injury.

J Neurotrauma 24:823–834

180. Boatright KM, Deis C, Denault JB, Sutherlin DP, Salvesen GS

(2004) Activation of caspases-8 and -10 by FLIP(L). Biochem J

382:651–657

181. O’Reilly LA, Ekert P, Harvey N et al (2002) Caspase-2 is not

required for thymocyte or neuronal apoptosis even thoughclea-

vage of caspase-2 is dependent on both Apaf-1 and caspase 9.

Cell Death Differ 9:832–841

182. Kim IR, Murakami K, Chen NJ, Saibil SD, Matysiak-Zablocki

E, Elford AR, Bonnard M, Benchimol S, Jurisicova A, Yeh WC,

Ohashi PS (2009) DNA damage- and stress-induced apoptosis

occurs independently of PIDD. Apoptosis 14:1039–1049

183. Ho LH, Read SH, Dorstyn L, Lambrusco L, Kumar S (2008)

Caspase-2 is required for cell death induced by cytoskeletal

disruption. Oncogene 27:3393–3404

184. Park MS, Kim BS, Devarajan P (2007) Hypoxia/reoxygenation

injury induces apoptosis of LLC-PK1 cells by activation of

caspase-2. Pediatr Nephrol 22:202–208

185. Gogvadze V, Orrenius S, Zhivotovsky B (2006) Multiple path-

ways of cytochrome c release from mitochondria in apoptosis.

Biochim Biophys Acta 1757:639–647

186. Baptiste-Okoh N, Barsotti AM, Prives C (2008) A role for

caspase 2 and PIDD in the process of p53-mediated apoptosis.

Proc Natl Acad Sci USA 105:1937–1942

187. Comelli M, Genero N, Mavelli I (2009) Caspase-independent

apoptosis in Friend’s erythroleukemia cells: role of mitochondrial

ATP synthesis impairment in relocation of apoptosis-inducing

factor and endonuclease G. J Bioenerg Biomembr 41:49–59

188. Parcellier A, Gurbuxani S, Schmitt E, Solary E, Garrido C

(2003) Heat shock proteins, cellular chaperones that modulate

mitochondrial cell death pathways. Biochem Biophys Res

Commun 304:505–512

189. Laudanski K, Wyczechowska D (2006) The distinctive role ofsmall heat shock proteins in oncogenesis. Arch Immunol Ther

Exp (Warsz) 54:103–111

190. Garrido C, Schmitt E, Cande C, Vahsen N, Parcellier A,

Kroemer G (2003) HSP27 and HSP70: potentially oncogenic

apoptosis inhibitors. Cell Cycle 2:579–584

191. Lin CY, Wu HY, Wang PL, Yuan CJ (2010) Mammalian Ste20-

like protein kinase 3 induces a caspase-independent apoptotic

pathway. Int J Biochem Cell Biol 42:98–105

192. Strauss G, Westhoff MA, Fischer-Posovszky P, Fulda S,

Schanbacher M, Eckhoff SM, Stahnke K, Vahsen N, Kroemer

G, Debatin KM (2008) 4-hydroperoxy-cyclophosphamide

mediates caspase-independent T-cell apoptosis involving oxi-

dative stress-induced nuclear relocation of mitochondrial apop-

togenic factors AIF and EndoG. Cell Death Differ 15:332–343

193. Satou T, Cummings BJ, Cotman CW (1995) Immunoreactivity

for Bcl-2 protein within neurons in the Alzheimer’s disease

brain increases with disease severity. Brain Re 697:35–43

56 Mol Cell Biochem (2011) 351:41–58

123

194. Jarskog LF, Gilmore JH (2000) Developmental expression of

Bcl-2 protein in human cortex. Brain Res Dev Brain Res

119:225–230

195. Levy OA, Malagelada C, Greene LA (2009) Cell death path-

ways in Parkinson’s disease: proximal triggers, distal effectors,

and final steps. Apoptosis 14:478–500

196. Morissette MR, Rosenzweig A (2005) Targeting survival sig-

naling in heart failure. Curr Opin Pharmacol 5:165–170

197. Donath MY, Halban PA (2004) Decreased beta-cell mass in

diabetes: significance, mechanisms and therapeutic implications.

Diabetologia 47:581–589

198. Kern TS, Du Y, Miller CM, Hatala DA, Levin LA (2010)

Overexpression of Bcl-2 in vascular endothelium inhibits the

microvascular lesions of diabetic retinopathy. Am J Pathol

176:2550–2558

199. Littlewood TD, Bennett MR (2003) Apoptotic cell death in

atherosclerosis. Curr Opin Lipidol 14:469–475

200. Susnow N, Zeng L, Margineantu D, Hockenbery DM (2009)

Bcl-2 family proteins as regulators of oxidative stress. Semin

Cancer Biol 9:42–49

201. Hajra KM, Liu JR (2004) Apoptosome dysfunction in human

cancer. Apoptosis 9:691–704

202. Yip KW, Reed JC (2008) Bcl-2 family proteins and cancer.

Oncogene 27:6398–6406

203. Reagan-Shaw S, Nihal M, Ahsan H, Mukhtar H, Ahmad N

(2008) Combination of vitamin E and delenium causes an

induction of apoptosis of human prostate cancer cells by

enhancing the Bax/Bcl-2 ration. Prostate 68:1624–1634

204. Cartron PF, Oliver L, Martin S, Moreau C, LeCabellec MT,

Jezequel P, Meflah K, Vallette FM (2002) The expression of a

new variant of the pro-apoptotic molecule Bax, Baxw, is cor-

related with an increased survival of glioblastoma multiforme

patients. Hum Mol Genet 11:675–687

205. Kirkin V, Joos S, Zornig M (2004) The role of Bcl-2 family

members in tumorigenesis. Biochim Biophys Acta 1644:229–249

206. O’Neill J, Manion M, Schwartz P, Hockenbery DM (2004)

Promises and challenges of targeting Bcl-2 anti-apoptotic pro-

teins for cancer therapy. Biochim Biophys Acta 1705:43–51

207. Juin P, Geneste O, Raimbaud E, Hickman JA (2004) Shooting at

survivors: Bcl-2 family members as drug targets for cancer.

Biochim Biophys Acta 1644:251–260

208. Frankel SR (2003) Oblimersen sodium (G3139 Bcl-2 antisense

oligonucleotide) therapy in Waldenstrom’s macroglobulinemia: a

targeted approach to enhance apoptosis. Semin Oncol 30:

300–304

209. Degterev A, Lugovskoy A, Cardone M, Mulley B, Wagner G,

Mitchison T, Yuan J (2001) Identification of small-molecule

inhibitors of interaction between the BH3 domain and Bcl-xL.

Nat Cell Biol 3:173–182

210. Tzung SP, Kim KM, Basanez G, Giedt CD, Simon J, Zimmer-

berg J, Zhang KY, Hockenbery DM (2001) Antimycin A mimics

a cell-death-inducing Bcl-2 homology domain 3. Nat Cell Biol

3:183–191

211. Enyedy IJ, Ling Y, Nacro K, Tomita Y, Wu X, Cao Y, Guo R,

Li B, Zhu X, Huang Y, Long YQ, Roller PP, Yang D,

Wang S (2001) Discovery of small-molecule inhibitors of Bcl-2

through structure-based computer screening. J Med Chem

44:4313–4324

212. Kitada S, Leone M, Sareth S, Zhai D, Reed JC, Pellecchia M

(2003) Discovery, characterization, and structure-activity rela-

tionships studies of proapoptotic polyphenols targeting B-cell

lymphocyte/leukemia-2 proteins. J Med Chem 46:4259–4264

213. Becattini B, Kitada S, Leone M, Monosov E, Chandler S, Zhai

D, Kipps TJ, Reed JC, Pellecchia M (2004) Rational design and

real time, in-cell detection of the proapoptotic activity of a novel

compound targeting Bcl-X(L). Chem Biol 11:389–395

214. Li J, Viallet J, Haura EB (2008) A small molecule pan-Bcl-2

family inhibitor, GX15–070, induces apoptosis and enhances

cisplatin-induced apoptosis in non-small cell lung cancer cells.

Cancer Chemother Pharmacol 61:525–534

215. McGregor N, Patel L, Craig M, Weidner S, Wang S, Pienta KJ

(2010) AT-101 (R-(-)-gossypol acetic acid) enhances the

effectiveness of androgen deprivation therapy in the VCaP

prostate cancer model. J Cell Biochem 110:1187–1194

216. Ferrer P, Asensi M, Priego S, Benlloch M, Mena S, Ortega A,

Obrador E, Esteve JM, Estrela JM (2007) Nitric oxide mediates

natural polyphenol-induced Bcl-2 down-regulation and activa-

tion of cell death in metastatic B16 melanoma. J Biol Chem

282:2880–2890

217. Muilenburg DJ, Coates JM, Virudachalam S, Bold RJ (2010)

Targeting Bcl-2-mediated cell death as a novel therapy in pan-

creatic cancer. J Surg Res 163(2):276–281

218. An J, Chen Y, Huang Z (2004) Critical upstream signals of

cytochrome C release induced by a novel Bcl-2 inhibitor. J Biol

Chem 279:19133–19140

219. Wang JL, Zhang ZJ, Choksi S, Shan S, Lu Z, Croce CM,

Alnemri ES, Korngold R, Huang Z (2000) Cell permeable Bcl-2

binding peptides: a chemical approach to apoptosis induction in

tumor cells. Cancer Res 60:1498–1502

220. Yin H, Lee GI, Sedey KA, Kutzki O, Park HS, Orner BP, Ernst

JT, Wang HG, Sebti SM, Hamilton AD (2005) Terphenyl-Based

Bak BH3 alpha-helical proteomimetics as low-molecular-weight

antagonists of Bcl-xL. J Am Chem Soc 127:10191–10196

221. Walensky LD, Kung AL, Escher I, Malia TJ, Barbuto S, Wright

RD, Wagner G, Verdine GL, Korsmeyer SJ (2004) Activation of

apoptosis in vivo by a hydrocarbon-stapled BH3 helix. Science

305:1466–1470

222. Bombrun A, Gerber P, Casi G, Terradillos O, Antonsson B,

Halazy S (2003) 3, 6-dibromocarbazole piperazine derivatives

of 2-propanol as first inhibitors of cytochrome c release via Bax

channel modulation. J Med Chem 46:4365–4368

223. Becattini B, Sareth S, Zhai D, Crowell KJ, Leone M, Reed JC,

Pellecchia M (2004) Targeting apoptosis via chemical design:

inhibition of bid-induced cell death by small organic molecules.

Chem Biol 11:1107–1117

224. Guo B, Zhai D, Cabezas E, Welsh K, Nouraini S, Satterthwait

AC, Reed JC (2003) Humanin peptide suppresses apoptosis by

interfering with Bax activation. Nature 423:456–461

225. Sawada M, Sun W, Hayes P, Leskov K, Boothman DA,

Matsuyama S (2003) Ku70 suppresses the apoptotic transloca-

tion of Bax to mitochondria. Nat Cell Biol 5:320–329

226. Foster FM, Owens TW, Tanianis-Hughes J, Clarke RB, Brennan

K, Bundred NJ, Streuli CH (2009) Targeting inhibitor of

apoptosis proteins in combination with ErbB antagonists in

breast cancer. Breast Cancer Res 1:R41

227. Liston P, Fong WG, Korneluk RG (2003) The inhibitors of

apoptosis: there is more to life than Bcl2. Oncogene 22:8568–

8580

228. LaCasse EC, Cherton-Horvat GG, Hewitt KE, Jerome LJ,

Morris SJ, Kandimalla ER, Yu D, Wang H, Wang W, Zhang R,

Agrawal S, Gillard JW, Durkin JP (2006) Preclinical charac-

terization of AEG35156/GEM 640, a second-generation anti-

sense oligonucleotide targeting X-linked inhibitor of apoptosis.

Clin Cancer Res 12:5231–5241

229. Wu TY, Wagner KW, Bursulaya B, Schultz PG, Deveraux QL

(2003) Development and characterization of nonpeptidic small

molecule inhibitors of the XIAP/caspase-3 interaction. Chem

Biol 10:759–767

230. Schimmer AD, Welsh K, Pinilla C, Wang Z, Krajewska M,

Bonneau MJ, Pedersen IM, Kitada S, Scott FL, Bailly-Maitre B,

Glinsky G, Scudiero D, Sausville E, Salvesen G, Nefzi A,

Ostresh JM, Houghten RA, Reed JC (2004) Small-molecule

Mol Cell Biochem (2011) 351:41–58 57

123

antagonists of apoptosis suppressor XIAP exhibit broad antitu-

mor activity. Cancer Cell 5:25–35

231. Chen J, Nikolovska-Coleska Z, Wang G, Qiu S, Wang S (2006)

Design, synthesis, and characterization of new embelin deriva-

tives as potent inhibitors of X-linked inhibitor of apoptosis

protein. Bioorg Med Chem Lett 16:5805–5808

232. Li L, Thomas RM, Suzuki H, De Brabander JK, Wang X,

Harran PG (2004) A small molecule Smac mimic potentiates

TRAIL- and TNFalpha-mediated cell death. Science 305:

1471–1474

233. Zobel K, Wang L, Varfolomeev E, Franklin MC, Elliott LO,

Wallweber HJ, Okawa DC, Flygare JA, Vucic D, Fairbrother

WJ, Deshayes K (2006) Design, synthesis, and biological

activity of a potent Smac mimetic that sensitizes cancer cells to

apoptosis by antagonizing IAPs. ACS Chem Biol 1:525–533

234. Oost TK, Sun C, Armstrong RC, Al-Assaad AS, Betz SF,

Deckwerth TL, Ding H, Elmore SW, Meadows RP, Olejniczak

ET, Oleksijew A, Oltersdorf T, Rosenberg SH, Shoemaker AR,

Tomaselli KJ, Zou H, Fesik SW (2004) Discovery of potent

antagonists of the antiapoptotic protein XIAP for the treatment

of cancer. J Med Chem 47:4417–4426

235. Lu J, Bai L, Sun H, Nikolovska-Coleska Z, McEachern D, Qiu

S, Miller RS, Yi H, Shangary S, Sun Y, Meagher JL, Stuckey

JA, Wang S (2008) SM-164: a novel, bivalent Smac mimetic

that induces apoptosis and tumor regression by concurrent

removal of the blockade of cIAP-1/2 and XIAP. Cancer Res

68:9384–9393

236. Chauhan D, Neri P, Velankar M, Podar K, Hideshima T, Fulc-

initi M, Tassone P, Raje N, Mitsiades C, Mitsiades N, Rich-

ardson P, Zawel L, Tran M, Munshi N, Anderson KC (2007)

Targeting mitochondrial factor Smac/DIABLO as therapy for

multiple myeloma (MM). Blood 109:1220–1227

237. Arnt CR, Chiorean MV, Heldebrant MP, Gores GJ, Kaufmann

SH (2002) Synthetic Smac/DIABLO peptides enhance the

effects of chemotherapeutic agents by binding XIAP and cIAP1

in situ. J Biol Chem 277:44236–44243

58 Mol Cell Biochem (2011) 351:41–58

123