ArticlePDF AvailableLiterature Review

Youle RJ, Strasser A.. The BCL-2 protein family: opposing activities that mediate cell death. Nat Rev Mol Cell Biol 9: 47-59

Authors:

Abstract

BCL-2 family proteins, which have either pro- or anti-apoptotic activities, have been studied intensively for the past decade owing to their importance in the regulation of apoptosis, tumorigenesis and cellular responses to anti-cancer therapy. They control the point of no return for clonogenic cell survival and thereby affect tumorigenesis and host-pathogen interactions and regulate animal development. Recent structural, phylogenetic and biological analyses, however, suggest the need for some reconsideration of the accepted organizational principles of the family and how the family members interact with one another during programmed cell death. Although these insights into interactions among BCL-2 family proteins reveal how these proteins are regulated, a unifying hypothesis for the mechanisms they use to activate caspases remains elusive.
The BCL-2 (B-cell lymphoma-2) gene was discovered
at the t(14;18) chromosome translocation breakpoint
in B-cell follicular lymphomas, where its transcription
becomes excessively driven by the immunoglobulin heavy
chain gene promoter and enhancer on chromosome 14
(REFS 1–3). One key early discovery that introduced a new
paradigm for carcinogenesis was that overexpression of
BCL-2 does not promote cell proliferation as most pre-
viously discovered oncogenes do; rather, overexpression
of BCL-2 inhibits cell death
4
. Apoptosis has now been
widely accepted as a prominent tumour-suppression
mechanism. Mutations in certain oncogenes that result
in the activation of cell proliferation, such as deregulated
MYC expression, require a second mutation to inhibit
the apoptosis machinery so that tumour promotion can
proceed efficiently
5,6
. Thus, the combined overexpression
of BCL-2 and MYC synergize potently in the develop-
ment of lymphomas and certain other types of cancer
7
. It
has also become clear that, beyond roles in cancer, BCL-2
and other members of the family are essential for an array
of apoptosis programmes, including developmentally
programmed cell death, tissue turnover and host defence
against pathogens.
In mammals, there are at least 12 core BCL-2 family
proteins, including BCL-2 itself and proteins that have
either three-dimensional (3D) structural similarity or a
predicted secondary structure that is similar to BCL-2
(FIG. 1). These proteins display a range of bioactivities,
from inhibition to promotion of apoptosis. Numerous
so-called BH3-only proteins share homology with each
other and the remainder of the BCL-2 protein family
only through the short BH3 motif
8
. Other than BID, the
predicted overall structures of the BH3-only proteins
seem to be unrelated and appear to lack a close evolution-
ary relationship to the core members of the BCL-2 family
9
.
But, all BH3-only proteins interact with and regulate
the core BCL-2 family proteins to promote apoptosis.
Several of the members of these two classes have been
knocked out in mice to reveal their physiological roles,
redundancy and interactions in vivo (TABLE 1).
This review covers recent insights into the biochemical,
cellular and physiological roles of the BCL-2 family with-
out reiterating the roles of these proteins in cancer and
drug development, which have recently been expertly
reviewed
5,10
. With the recent advances in understanding
BCL-2 family protein interactions, we focus on how such
interactions lead these proteins to change subcellular local-
ization and conformation to regulate their bioactivities.
The latest progress into the differential regulation of
organ development, maintenance and tissue turnover in
mice by BCL-2 family members are also reviewed.
BCL-2 family proteins in apoptosis
All pathways to apoptosis converge on the activation of
caspases, which are cysteinyl aspartate proteases that
coordinate the efficient dismantling and engulfment
of doomed cells (FIG. 2). Two pathways of cell death can
be distinguished by whether they require BCL-2 family
proteins and by which caspases are crucial for their exe-
cution. The intrinsic pathway — also called the BCL-2-
regulated or mitochondrial pathway (in reference to
the role these organelles play) — is activated by various
*Biochemistry Section,
Surgical Neurology Branch,
National Institute of
Neurological Disorders and
Stroke, The National
Institutes of Health,
Bethesda, Maryland 20892,
USA.
The Walter and Eliza Hall
Institute of Medical Research,
Parkville 3050,
Melbourne, Australia.
e-mails:
youler@ninds.nih.gov;
strasser@wehi.edu.au
doi:10.1038/nrm2308
BH3 motif
The amino-acid sequence
LXXXGD, in which X represents
any amino acid. This motif is
conserved between most core
BCL-2 family members and
among BH3-only proteins.
The BCL-2 protein family: opposing
activities that mediate cell death
Richard J. Youle* and Andreas Strasser
Abstract | BCL-2 family proteins, which have either pro- or anti-apoptotic activities, have
been studied intensively for the past decade owing to their importance in the regulation of
apoptosis, tumorigenesis and cellular responses to anti-cancer therapy. They control the
point of no return for clonogenic cell survival and thereby affect tumorigenesis and host
pathogen interactions and regulate animal development. Recent structural, phylogenetic
and biological analyses, however, suggest the need for some reconsideration of the accepted
organizational principles of the family and how the family members interact with one
another during programmed cell death. Although these insights into interactions among
BCL-2 family proteins reveal how these proteins are regulated, a unifying hypothesis for the
mechanisms they use to activate caspases remains elusive.
REVIEWS
NATURE REVIEWS
|
MOLECULAR CELL BIOLOGY VOLUME 9
|
JANUARY 2008
|
47
© 2008 Nature Publishing Group
3
BIM/BOD
3
BAD
3
BCL-G
BID (2BID)
PUMA/BBC3
BCL-XL (1R2D)
BCL-W (1MK3)
MCL1
BAX (1F16)
BAK (2IMS)
BIK/BLK/NBK
HRK/DP5
WWE
UBA
HECT
MULE
A1
Boo/Diva/BCL-B/BCL2L10
BCL-RAMBO
BOK/MTD
NOXA
BMF
BCL-2 (1G5M)
BCL
-2 homo
l
ogue
s
BH3-onl
y
Anti-apoptotic
Pro-apoptotic
Nature Reviews | Molecular Cell Biology
BH1
BH2
BH3
BH4
TM region
α-helix
UBA
WWE
HECT
TNF receptor family
Cell-surface receptors in the
tumour necrosis factor (TNF)
family.
developmental cues or cytotoxic insults, such as viral
infection, DNA damage and growth-factor deprivation,
and is strictly controlled by the BCL-2 family of proteins.
This pathway predominantly leads to the activation
of caspase-9 (REF. 11) but, at least in certain cell types,
the intrinsic pathway can proceed in the absence of
caspase-9 or its activator, apoptotic protease-activating
factor-1 (APAF1)
12
.
The extrinsic or death-receptor pathway is triggered
by ligation of so-called death receptors (members of
the tumour necrosis factor (TNF) receptor family, such as
Fas or TNF receptor-1 (TNFR1)) that contain an intra-
cellular death domain, which can recruit and activate
caspase-8 through the adaptor protein Fas-associated
death domain (FADD; also known as MORT1) at the cell
surface. This recruitment causes subsequent activation of
Figure 1 | Sequence alignment of core BCL-2 family proteins and BH3-only proteins. Green bars depict α-helical
segments from the determined structures (when labelled by Protein Data Bank (PDB) identifier in parentheses) or from
secondary structure prediction (as predicted using PSIPRED). Red lines label regions of predicted transmembrane (TM)
domains (as predicted using TMHMM). Sequence homologies of the BH1 (brown lines), BH2 (grey lines), BH3 (blue lines) and
BH4 (orange lines) regions are shown. The BH1, BH2 and BH3 domains fold to line a hydrophobic pocket that can bind BH3-
only peptides. The BH3 domain, particularly among the BH3-only proteins, mediates interaction between the BH3-only
proteins and core BCL-2 family proteins and thereby promotes apoptosis. The upper five proteins (BCL-2, BCL-XL, BCL-W,
A1 and MCL1) are generally anti-apoptotic. The three proteins in the shaded area are less well studied and cannot be
categorized at this time. The lower 12 proteins are considered to be pro-apoptotic. MULE contains a ubiquitin-associated
domain (UBA), the Trp-Trp-Glu interaction module (WWE) and a HECT ubiquitin ligase domain. BID has a unique role as both
a BCL-2 homologue and a BH3-only protein and links the intrinsic and extrinsic apoptosis pathways (FIG. 2). BIM (also known
as BOD), BAD and BMF are unstructured proteins.
REVIEWS
48
|
JANUARY 2008
|
VOLUME 9 www.nature.com/reviews/molcellbio
© 2008 Nature Publishing Group
Death domain
A protein-interaction module
that consists of six α-helices
and that is involved in
apoptosis and other signalling
pathways.
downstream (effector) caspases, such as caspase-3, -6 or
-7, without any involvement of the BCL-2 family. In some
cells, most notably hepatocytes, the extrinsic pathway
can intersect the intrinsic pathway through caspase-8
cleavage-mediated activation of the pro-apoptotic BH3-
only protein BID
13,14
. The C-terminal truncated form of
BID (tBID) translocates to mitochondria and promotes
further caspase activation (caspase-9 and the effector
caspases caspase-3, -6 and -7) through the intrinsic path-
way. In these situations, loss of BID or overexpression
of BCL-XL inhibits cell death
13
.
BCL-2 family proteins have opposing apoptotic activities.
BCL-2 family members have classically been grouped
into three classes. One class inhibits apoptosis (BCL-2,
BCL-XL, BCL-W, MCL1, BCL-B (also known as
BCL-2L10) and A1 (also known as BCL-2A1), whereas
a second class promotes apoptosis (BAX, BAK and
BOK (also known as MTD)). A third divergent class
of BH3-only proteins (BAD, BIK (also known as BLK
or NBK), BID, HRK (also known as death protein-5
(DP5)), BIM (also known as BOD), BMF, NOXA and
PUMA (also known as BBC3)) have a conserved BH3
domain that can bind and regulate the anti-apoptotic
BCL-2 proteins to promote apoptosis (FIG. 1). It appears
that the pro-apoptotic family members BAX and BAK
are crucial for inducing permeabilization of the outer
mitochondrial membrane (OMM) and the subsequent
release of apoptogenic molecules (such as cytochrome c
and DIABLO (also known as SMAC)), which leads to
caspase activation. The anti-apoptotic family members,
such as BCL-2 and BCL-XL, inhibit BAX and BAK.
Recent evidence indicates that BH3-only proteins de-
repress BAX and BAK by direct binding and inhibition
of BCL-2 and other anti-apoptotic family members
15
.
By contrast, an opposing model postulates direct activ-
ation of BAX and BAK by some BH3-only proteins
(specifically BIM, tBID and PUMA)
16
(FIG. 2).
Table 1 | Phenotypes of mice that are deficient in BCL-2 family members
BCL-2 family
member
Defects caused by its deletion* Refs
Pro-survival family members
BCL-2 Abnormal death of renal epithelial progenitors, melanocyte progenitors and mature B
and T lymphocytes. Causes fatal polycystic kidney disease (100% mortality by 6 weeks),
premature greying and lymphopoenia (but all of these effects can be rescued by
concomitant loss of the BH3-only protein BIM).
130
BCL-XL Abnormal death of fetal erythroid progenitors and neuronal cells. Causes death around
embryonic day 14 (100% mortality).
129
BCL-W Abnormal death of developing sperm cells. Causes male sterility. 132
A1A Abnormally accelerated death of granulocytes and mast cells in culture. 133
MCL1 Failure in implantation. Conditional knockout causes premature death of immature and
mature B and T lymphoid cells, as well as haemopoietic stem cells.
128
Pro-apoptotic BAX/BAK family members
BAX Mild lymphoid hyperplasia, male sterility due to sperm-cell differentiation defect. 135
BAK No obvious defects detected so far. 136
Pro-apoptotic BH3-only proteins
BIM Lymphoid and myeloid cell hyperplasia, fatal SLE-like autoimmune disease (on mixed
genetic C57BL/6x129SV background), many cell types are abnormally resistant to
cytokine deprivation, deregulated calcium flux and the chemotherapeutic drug taxol;
mild but significant resistance of many cell types to DNA damage and glucocorticoids.
143
BID BID-deficient mice are resistant to Fas-activation-induced hepatocyte killing and fatal
hepatitis; however, some cell types (such as lymphoid cells) are normally sensitive to
Fas-induced apoptosis.
13, 14
PUMA Many cell types are profoundly resistant to DNA damage; many are also resistant to
cytokine deprivation, glucocorticoids and phorbol ester.
150,151
BAD Mild resistance of some cell types to deprivation of epidermal growth factor or insulin
growth factor.
154
HRK Abnormal, although relatively mild, resistance of certain neuronal populations to
deprivation of nerve growth factor.
155,156
BIK No obvious defects detected so far. 158
NOXA
Relatively mild resistance of fibroblasts to γ-irradiation or etoposide, but profound
resistance of these same cells and keratinocytes in the skin to ultraviolet irradiation.
150
*These are phenotypes found in mice. The roles of these proteins may differ in humans. BAD, BCL-2 antagonist of cell death;
BAK, BCL-2-antagonist/killer-1; BAX, BCL-2-associated X protein; BCL-2, B-cell lymphoma-2; A1A, BCL-2-related protein A1A;
BCL-W, BCL-2-like-2; BCL-XL, a BCL-2-like protein; BID, BH3-interacting domain death agonist; BIK, BCL-2-interacting killer;
BIM, BCL-2-like-11; HRK, harakiri (also known as death protein-5); MCL1, myeloid cell leukaemia sequence-1; PUMA, BCL-2
binding component-3; SLE, systemic lupus erythematosus.
REVIEWS
NATURE REVIEWS
|
MOLECULAR CELL BIOLOGY VOLUME 9
|
JANUARY 2008
|
49
© 2008 Nature Publishing Group
Nature Reviews | Molecular Cell Biolog
y
Growth-factor deprivation,
stress, UV, viruses
Inactive
BH3-only
BCL-2
BAX/BAK
Cytochrome c release,
mitochondrial fragmentation
APAF1 assembly
into apoptosome
FAS
Caspase-8
TNFR1
APAF1
Cytochrome c
BID cleavage
Intrinsic pathway Extrinsic pathway
Active
BH3-only
Caspase-3
Apoptosis
Mitochondrial outer
membrane permeabilization
The process by which the outer
membrane of mitochondria
leaks certain soluble
intermembrane space proteins,
such as cytochrome c, into the
cytoplasm.
Apoptosome
The caspase-9 activation
complex that is composed of
APAF1 heptamers and that is
assembled on binding of
APAF1 monomers to
cytochrome c.
BAX and BAK promote caspase activation by their
effects on mitochondria. Either directly or indirectly,
these two pro-apoptotic BCL-2 family members induce
the release of proteins from the space between the inner
and outer mitochondrial membranes
17
. This process of
mitochondrial outer membrane permeabilization (MOMP)
results in the release of cytochrome c and other soluble
proteins into the cytosol. Although it is commonly
thought that BAX and BAK form pores in membranes, the
biochemical nature of such pores and how anti-apoptotic
BCL-2 family proteins might regulate them remains a
key and controversial issue in the field of cell death
18
. At
the same time as cytochrome c release (or immediately
before), BAX and BAK induce mitochondria to fragment
into more numerous and smaller units, which suggests
connections between mitochondrial division processes
and the functions of the BCL-2 family
19
.
Once the OMM has been permeabilized, soluble
proteins diffuse from the intermembrane space into the
cytosol, where they promote caspase activation. The best
studied of these proteins is cytochrome c, which binds to
APAF1 and leads to the assembly of a heptameric protein
ring called an apoptosome, which can bind pro-caspase-9
and induce its activation through a conformational
change
20,21
. Cytochrome c–APAF1-dependent activa-
tion of caspase-9 is absolutely required for neuronal and
fibroblast cell-death processes
22
. However, in addition to
this process, lymphocytes can probably use alternative
APAF1-, caspase-9- and cytochrome c-independent,
but pro-apoptotic BCL-2-family-member-dependent,
pathways for caspase activation and cell killing
12,22
.
Intriguingly, caspase activation in lymphocytes can be
amplified by APAF1 even when APAF1 has not been
incorporated into the apoptosome
22
.
One APAF1-independent pathway of caspase
activation is the relief of caspase inhibition by inhibitor
of apoptosis proteins (IAPs), such as XIAP, which bind
and neutralize certain caspases (such as caspase-9 and
caspase-3). This inhibitory action of IAPs can be antago-
nized by the binding of DIABLO, which is released
from mitochondria after the activation of BAX and/or
BAK. However, DIABLO-deficient mice
23
, as well as
XIAP-deficient mice
24
, do not display significant apop-
totic phenotypes, which suggests that novel processes
of caspase activation remain to be discovered. Several
APAF1 related proteins, called NOD-like receptors, regulate
alternative pathways of caspase activation that occur in
non-apoptotic host defence processes that are associated
with innate immunity and serve as examples of pathways
that can also have roles during apoptosis
25
. One of these
NOD-like receptors, NALP1, can be regulated by BCL-2
and BCL-XL
26
in manner that is reminiscent of caspase
activation in the worm (BOX 1).
BCL-2 and BCL-XL appear to control cell survival
beyond the APAF1–caspase-9 axis. If caspase activation
is inhibited by loss of APAF1 or caspase-9, or even by the
combined loss of caspase-9 and caspase-2, the rate of acqui-
sition of apoptotic morphology of myeloid progenitors
and mast cells induced by growth-factor withdrawal or
DNA damage can be significantly delayed. However,
although the onset of apoptotic morphology can be
delayed, the cells still lose clonogenic potential and thus
effectively die, unlike cells that overexpress BCL-2 or
BCL-XL
27,28
. Thus, the step of apoptosis regulation that is
controlled by the BCL-2 family appears to be the most gen-
eral final commitment step for the decision between cell
life and death. The disruption of mitochondria by BAX and
BAK may be one cause of eventual clonogenic cell death in
the absence of apoptosome activation. Normally, caspase
activation rapidly and efficiently mediates cell demoli-
tion and removal. When caspases are blocked, certain
features of apoptosis can be lost (or delayed), which causes
the cells to die more slowly by BCL-2-family-mediated
mitochondrial disruption or by novel caspase-activation
pathways that have yet to be characterized.
Figure 2 | Scheme depicting intrinsic and extrinsic pathways of apoptosis. Apoptosis
can be induced by cell surface receptors, such as Fas and tumour necrosis factor
receptor-1 (TNFR1) (extrinsic pathway, right), or by various genotoxic agents, metabolic
insults or transcriptional cues (intrinsic pathway, left). The intrinsic pathway starts with
BH3-only protein induction or post-translational activation, which results in the
inactivation of some BCL-2 family members. This relieves inhibition of BAX and BAK
activation, which in turn promotes apoptosis. Some BH3-only proteins, such as BIM and
PUMA, may also be able to activate BAX and/or BAK (as shown by the dotted line). Once
activated, BAX and BAK promote cytochrome c release and mitochondrial fission, which
leads to the activation of APAF1 into an apoptosome and activates caspase-9 to activate
caspase-3. Caspases in turn cleave a series of substrates, activate DNases and orchestrate
the demolition of the cell. The extrinsic pathway can bypass the mitochondrial step and
activate caspase-8 directly, which leads to caspase-3 activation and cell demolition. The
BCL-2 family regulates the intrinsic pathway and can modulate the extrinsic pathway
when cleavage of BID communicates between the two pathways.
REVIEWS
50
|
JANUARY 2008
|
VOLUME 9 www.nature.com/reviews/molcellbio
© 2008 Nature Publishing Group
Nature Reviews | Molecular Cell Biology
EGL-1
(BH3-only)
CED-9
(core BCL-2 family)
CED-4
(APAF1-like)
CED-3
(caspase)
Inhibitor of apoptosis
protein
(IAP). One of a family of
proteins that inhibits apoptosis
by binding or degrading
caspases.
NOD-like receptor
A cytosolic receptor that is
homologous to NOD1 and is
involved in innate immunity
pathways.
E3 ligase
One of a family of proteins that
facilitate the transfer of
ubiquitin from a donor protein
to a specific substrate protein
that may signal the target for
proteosomal degradation.
Structure and evolution
The core multi-BH domain BCL-2 family members and,
surprisingly, the BH3-only protein BID (FIG. 1) have con-
served regions of sequence homology and similar pre-
dicted secondary structure. Structures of seven of these
proteins (BCL-XL
29
, BCL-2 (REF. 30), BCL-W
31,32
, MCL1
(REF. 33), BAX
34
, BAK
35
and BID
36,37
) show remarkable
similarity, which is intriguing considering that some are
pro-apoptotic and others are anti-apoptotic. Ks-BCL-2,
a viral homologue of BCL-2 (REF. 38), as well as two viral
proteins without apparent sequence similarity to BCL-2
family proteins, M11L
39,40
and N1L
41
, display a helical
fold that is similar to that of BCL-XL, and these inhibit
apoptosis, which indicates that several viruses use BCL-2
family members to counteract host defence.
The 3D structures of the seven core BCL-2 family
proteins mentioned above have yet to reveal any distin-
guishing difference between anti-apoptotic members
(such as BCL-XL and MCL1) and pro-apoptotic mem-
bers (such as BAX and BID). All seven proteins are helical
bundles with a hydrophobic helix-turn-helix hairpin that
is flanked on both sides by pairs of amphipathic helices.
Excluding the viral anti-apoptotic BCL-2-like proteins,
BCL-2 homologues appear to have C-terminal membrane-
anchoring domains. In addition, pro-apoptotic BID appears
to be myristoylated to mediate membrane anchorage
42
.
In three proteins, BAX, BCL-W and MCL1, the C-terminal
anchor has been included in the structural analysis and
fits into a hydrophobic pocket formed by the BH1, -2
and -3 regions. The same pocket that sequesters the
C-terminal membrane anchor can also bind to peptides
of the BH3-domain sequences of BAK, BAD and BIM
43–45
,
which suggests that it also functions in dimerization with
BH3-only proteins and/or multi-BH-domain-containing
BCL-2 family members (FIG. 3). An extended BIM BH3
peptide that is 23 amino acids in length binds along the
hydrophobic groove of BCL-XL, although it is inverted in
the C- to N-terminal helical direction relative to the ori-
entation of the BAX and BCL-W C-terminal membrane
anchors
45
(FIG. 3).
Classically, BH3-only proteins have been defined as
having homology to the core BCL-2 family members in
only the BH3 domain. Recent sequence analyses indicate
that, except for BID, the BH3-only proteins have pre-
dicted secondary structures or determined 3D structures
that are unrelated to the core BCL-2 family members,
and, except for BID, they probably acquired BH3 motifs
by convergent evolution
9
. One particular example of
a BH3-motif-containing protein that is not otherwise
related to the core BCL-2 family is MULE, an E3 ligase
that reportedly targets MCL1 for ubiquitylation and pro-
teasomal degradation. MULE has a BH3 domain and can
loosely be considered to belong to the class of BH3-only
proteins, which interact with and regulate other mem-
bers of the BCL-2 family
46,47
(FIG. 4). Even the autophagy
regulatory protein beclin-1 reportedly binds to BCL-2
through a BH3 domain, although further biochemical
and genetic experiments are needed to establish a func-
tional connection
48
. BAD, BMF and BIM are intrinsically
unstructured
49
and, along with PUMA, these proteins
are not likely to be core BCL-2 family homologues on the
basis of secondary structure predictions (FIG. 1).
So, BH3-only proteins include various proteins that
share a single motif that allows them to bind and regu-
late the core BCL-2 family members. BID, by contrast, is
the one BH3-only protein with a determined structure
that places it squarely in the core BCL-2 family mem-
bers, which perhaps explains why BID shares certain
properties with multidomain BCL-2 family members,
such as the ability to oligomerize
50
and to permeabilize
membranes
51
. Thorough phylogenetic analyses of the
BCL-2 family have generated important insights into
the origins of the core BCL-2 family members (BOX 2)
and the BH3-only proteins, and suggest that many of
these proteins might have biological activities beyond
regulation of cell death
9,52
.
BCL-2 family protein activation
BH3-only proteins are pro-apoptotic and function as
initial sensors of apoptotic signals that emanate from
various cellular processes. BH3-only protein expression
can be induced by transcription factors. For example,
NOXA and PUMA are induced by the tumour sup-
pressor p53 in response to DNA damage
53–55
, and BIM
is induced by the class O forkhead box transcription
factor-3A (FOXO3A) in response to growth-factor
deprivation
56
and by the transcription factors CEBPα
Box 1 | The mechanism of CED-9, the C. elegans orthologue of BCL-2
Genetic analyses of the apoptosis pathway in Caenorhabditis elegans and recent
biochemical insights are consistent with the model shown in the figure above. EGL‑1, a
BH3‑only protein, is transcriptionally induced by developmental cues for programmed
cell death. EGL‑1 binds to the BCL‑2 homologue CED‑9, thereby freeing CED‑4, an
AAA+ ATPase that is related to apoptotic protease‑activating factor‑1 (APAF1), which is
normally sequestered by CED‑9. The released CED‑4 assembles into a tetrameric
apoptosome and activates the protease activity of the caspase CED‑3. This model differs
from the cytochrome c release model of mammalian cells (FIG. 2). It is not anticipated
that homologous proteins regulate the same process by different mechanisms, so some
underlying common process among BCL‑2 family members of C. elegans and mammals
may await discovery.
Two studies indicate that CED‑9 may do more than prevent CED‑4 activation. In
worms with mild ced‑3 loss‑of‑function mutations in which some excess cells survive,
weak loss‑of‑function ced‑9 alleles actually increase cell survival, which suggests that
CED‑9 also has pro‑apoptotic activity
163
. This might indicate that, depending on its
conformation, CED‑9 can have BCL‑2‑like (that is, anti‑apoptotic) or BAX‑like (that is,
pro‑apoptotic) activity
94
. In addition, loss of CED‑9 activity inhibits cell death due to
overexpression of drp‑1 (REF. 122), which further suggests that the sole core BCL‑2
family protein in C. elegans can function in both pro‑ and anti‑apoptotic modes.
Certain mammalian BCL‑2 family members have also been reported to be convertible
between anti‑ and pro‑apoptotic forms
164,165
. CED‑9 resides on mitochondria, as many
mammalian BCL‑2 family proteins do, but how this localization relates to its
biochemical action remains unclear. Similar to programmed cell death in mammals and
flies, mitochondria become fragmented during apoptosis in the worm
122
upstream of
caspase activation, showing that there is one common denominator involving
mitochondria in all three systems.
Future studies should explore how BCL‑2 family members function in sponges,
echinoderms and insects. One recent study in D. melanogaster came to the surprising
conclusion that although the two BCL‑2 family members are required for certain stress‑
induced apoptosis pathways, they are not required for developmentally programmed
cell death
166
.
REVIEWS
NATURE REVIEWS
|
MOLECULAR CELL BIOLOGY VOLUME 9
|
JANUARY 2008
|
51
© 2008 Nature Publishing Group
Nature Reviews | Molecular Cell Biology
a
b
c
d
e
f
ER stress
The accumulation of unfolded
or incompletely glycosylated
proteins in the endoplasmic
reticulum (ER) results in stress
that may lead to apoptosis.
Dynein motor complex
A molecular machine that
transports cargo along
microtubules.
JAK–STAT pathway
The Janus kinase (JAK)–signal
transducer and activator of
transcription (STAT) pathway is
a signalling pathway that is
activated by growth factors and
cytokines.
(CCAAT-enhancer binding protein-α) or CHOP (CEBP
homologous protein) in response to endoplasmic reticu-
lum (ER) stress
57
. BH3-only proteins can also be activated
post-translationally; for example, BAD is activated by
loss of phosphorylation in response to growth-factor
deprivation
58
; BID is activated by caspase-8-mediated
proteolysis
59,60
; BIM is activated by release from the
dynein motor complex
61
or by loss of extracellular signal-
regulated kinase (ERK)-mediated phosphorylation
(which targets it for ubiquitylation and proteasomal
degradation in healthy cells)
62,63
; BMF is activated by
release from actin–myosin motor complexes
64
; and BIK
is activated by an unknown mechanism in response to
inhibition of protein synthesis
65
.
Regulation of the expression levels of anti-apoptotic
BCL-2 family proteins is another way in which cells
can regulate apoptosis. For example, BCL-XL can be
transcriptionally induced by growth factors through the
Janus kinase–signal transducer and activator of trans-
cription (JAK–STAT) pathway to promote cell survival
66
.
MCL1 is rapidly degraded by the ubiquitin–proteasome
pathway in response to cytokine deprivation or other
death stimuli (such as ultraviolet (UV) radiation) and
can be upregulated post-transcriptionally to prevent
apoptosis by inhibiting the rate of degradation
46,67
.
Regulation of the expression levels of the pro-apoptotic
proteins BAX and BAK is less apparent and the pro-
teins appear to be constitutively expressed at more
or less constant levels. BAX and BAK are primarily
post-translationally regulated by other members of the
BCL-2 family.
When BH3-only proteins are induced or activated,
they interact with core BCL-2 family proteins to pro-
mote apoptosis. The binding of BH3-only proteins or
BH3 peptides to specific anti- and pro-apoptotic BCL-2
family members has been determined by using yeast
two-hybrid analysis, plasmon resonance binding assays
and by cell-free mitochondria and liposome permea-
bilization studies
68–72,73
. Together, these assays indicate
that some BH3-only proteins, such as BIM and PUMA,
bind all anti-apoptotic BCL-2 family members, whereas
others, such as BAD and NOXA, bind only certain anti-
apoptotic BCL-2 family members (FIG. 4). In addition to
interaction with anti-apoptotic BCL-2 family members,
several reports show synergy of BID or BIM with BAX
in cell-free membrane permeabilization assays, which
suggests that some BH3-only proteins may directly
bind and activate BAX and BAK
70–72
. However, it is dif-
ficult to detect binding of full-length BID, tBID or BIM
to BAX or BAK
15
, although a modified BH3 peptide can
bind BAX and/or BAK
74
. Other models in which BH3-
only proteins directly activate BAX and BAK are called
into question owing to results from Bim/Bid double
knockout mice and their cell lines, which show that
these putative direct activators of BAX and BAK are not
required for many apoptotic pathways
15
. Thus, known
BH3-only proteins appear to induce apoptosis primarily
by inhibiting anti-apoptotic BCL-2 family members,
thereby liberating BAX and BAK to cause MOMP and
activation of the caspase cascade (FIG. 2). The precise
biochemical mechanisms that lead to the activation
of BAX and BAK remain a mystery and constitute
the ‘holy grail’ of apoptosis research.
Dynamics of subcellular localization
The anti-apoptotic BCL-2 protein is embedded in the ER,
the nuclear envelope and the OMM by a hydrophobic
C-terminal membrane-spanning domain, with most of
its amino acids in the cytosol
75,76
. Although BCL-2 in any
of these subcellular locations can block apoptosis, the
functions of BCL-2 at the ER and the nuclear envelope
are less clear than those on mitochondria and have
recently been reviewed
77,78
.
In contrast to BCL-2, BAX is mostly cytosolic and
sequesters its hydrophobic C-terminal membrane anchor
in its BH3-binding pocket (FIG. 3), with a minor fraction
lightly bound to the OMM
79
. BAX appears to exist as a
monomer in the cytosol of cells rather than being bound
to any anti-apoptotic BCL-2 family members
80
. During
apoptosis induction, BAX translocates specifically to mito-
chondria (see Supplementary information S1 (movie)),
Figure 3 | Space-filling models of the structures of BAX, BCL-W and BCL-XL
bound to a BIM BH3-region peptide. Comparing the structures of full-length BCL-2
family members with those bound to BH3 peptides from other BCL-2 family members
suggests how subcellular localization might be linked to protein–protein interactions
among family members. (The structures in the lower panels are rotated 90 degrees from
those in the top panels.) Full-length BAX (a,b) and BCL-W (c,d) fold with the C-terminal
helix (grey) sequestered in a hydrophobic pocket. This C-terminal domain is
experimentally deleted in BCL-XL (e,f), which binds to an extended BH3 domain peptide
of BIM (yellow) in the homologous pocket that is occupied by the membrane anchor in
BAX, BCL-W and MCL1. The BIM peptide orientates in the pocket in the opposite
direction to the endogenous regions of the C terminus (that is, in the C- to N-terminal
direction). Because the C-terminal helix is involved in membrane binding and is thought
to penetrate deeply into membranes (FIG. 5), this helix would become displaced from the
hydrophobic pocket on mitochondrial translocation (see Supplementary information S1
(movie)). Emptying this pocket of the C-terminal helix would enable it to bind BH3
domains (yellow) from other BCL-2 family members, allowing hetero- or homodimer
formation. Alternatively, binding of BH3-only proteins to BAX or BCL-W in the cytosol
could displace the C-terminal helix from the pocket and trigger mitochondrial
translocation. Non-structured amino acids in BAX (1–12) and BCL-W (1–8 and 170–178)
have been excluded from the models.
REVIEWS
52
|
JANUARY 2008
|
VOLUME 9 www.nature.com/reviews/molcellbio
© 2008 Nature Publishing Group
Nature Reviews | Molecular Cell Biology
BCL-2
BCL-XL
MCL1
A1
BCL-W
PUMA
tBID
BIM
NOXA
BAD
3
3
where it inserts into the OMM as an integral membrane
protein
81
using its C-terminal membrane anchor
82
,
perhaps with organelle targeting specified by defined
regions in the N terminus
83
. This translocation step of
BAX, although reversible in certain situations, usually
correlates closely with the irreversible commitment
of cells to die and to the cytochrome c release step
discussed below.
BOK also translocates from the cytosol to mito-
chondria during apoptosis
84
, whereas BAK already
resides on the OMM (see subcellular localization of
BCL-2 family members in Supplementary information
S2 (table)) in healthy cells, where it has been reported
to be bound to MCL1 (REF. 67) and to BCL-XL
73
.
Notably, BAK–MCL1 and BAK–BCL-XL interaction
experiments rely on detergent extraction of membrane
proteins that, in some cases, can cause artefactual
interactions among BCL-2 family proteins
85
. Although
the OMM channel protein VDAC2 (voltage-dependent
anion channel-2) also reportedly binds to BAK
86
and
is important for BAK import into the OMM
87
, Vdac1/
Vdac2/Vdac3 triple knockout mice display normal
apoptosis
88
, indicating little or no role for VDACs in
BCL-2 family protein regulation. The mitochondrial
localization of BAK may be a consequence of MCL1
binding into the BH3 pocket of BAK and displacing
the C-terminal membrane anchor, allowing it to
interact with membranes. Although the BH3 domain
is the obvious candidate domain of MCL1 for this inter-
action, it is not exposed in the soluble MCL1 struc-
ture
33
. The C-terminal membrane anchor of BCL-XL
has been proposed to mediate binding to BAX by
fitting into its BH3-binding pocket in trans
89
, which
suggests another way by which BCL-XL and MCL1 can
interact with BAK. On apoptosis induction, MCL1 is
degraded (at least in certain cell types in response to
certain cytotoxic stimuli)
67,90
and/or the MCL1–BAK
and BCL-XL–BAK interactions are disrupted by BH3-
only proteins, such as NOXA, BIM
73
or BIK
65
, which
frees BAK to promote apoptosis.
Because the translocation of BAX to the mitochon-
dria correlates with pro-apoptotic activity, it is curi-
ous that BCL-2 is constitutively membrane bound but
BCL-XL
79,91
, BCL-W
92
and MCL1 (REF. 90) exist partly
in the cytosol and translocate from the cytosol to mito-
chondria during apoptosis. Their binding to BH3-only
proteins appears to be the trigger, perhaps by displacing
the C-terminal membrane anchors of BCL-XL
89
and
BCL-W
92
by occupancy of the hydrophobic pockets.
The intracellular translocation probably correlates
with conformational changes and deep insertion of
BCL-XL and BCL-W into the OMM. Whether this
translocation of anti-apoptotic BCL-2 family members
represents a mechanism to inhibit apoptosis, inactiva-
tion by BH3-only proteins or even conversion into
pro-apoptotic effectors remains unclear
92,93
. It has, for
example, been postulated that binding of a BH3-only
protein changes the conformation of the pro-survival
BCL-2-like protein so that it can then initiate formation
of BAX and/or BAK oligomers in the mitochondrial
and other intracellular membranes to cause initiator
caspase activation
94
.
Figure 4 | BH3-only protein binding specificity for BCL-2 homologues. BIM and PUMA bind to all five anti-apoptotic
BCL-2 family members tested. By contrast, NOXA only binds to MCL1 and A1, and BAD binds selectively to BCL-W, BCL-2
and BCL-XL. tBID binds avidly to BCL-XL, BCL-W, MCL1 and A1, but only weakly to BCL-2. BH3-only proteins do not
appear to bind strongly to BAX or BAK. These binding specificities recapitulate the ability of these proteins to activate
apoptosis. For example, BIM, BID or PUMA alone can induce apoptosis, whereas a combination of NOXA and BAD is
required. This probably enables the fine specificity of apoptosis regulation in different tissues and during changes in
cellular developmental stages. The BH3 domain is shown in red and the five anti-apoptotic BCL-2 family proteins are
shown in the middle of the figure.
Box 2 | Phylogenomics of BCL-2 family proteins
BCL‑2 gene orthologues have been identified in all metozoan animals examined so far
9
.
The earliest metazoan that has been analysed, the sponge
52
, contains two BCL‑2 related
genes that most closely resemble mammalian BOK
8
, which has so far received little
attention in mouse and human systems. Interestingly, Monosiga brevicollis, a single cell
choanoflagellate that is closely antecedent to sponges, has no identifiable BCL‑2 family
member in its recently sequenced genome (C. Wang, personal communication), which
suggests that BCL‑2 family genes evolved with multicellular life forms. However, some
recent viral gene products have been found with the signature helical fold of BCL‑2, and
these function in apoptosis regulation
39–41
, which suggests that additional structural
orthologues of BCL‑2 might exist in eukaryotes without discernible primary sequence
homology. The number of apparent BCL‑2 family member genes in different orders
varies widely. The sea urchin (Strongylocentrotus purpuratus) has ten homologous genes,
significantly more than insects (Drosophila melanogaster; two core BCL‑2 family
members) and round worms (Caenorhabditis elegans; one core BCL‑2 family member)
(BOX 1). C. elegans also has one prominent BH3‑only gene, egl‑1, which appears to be
crucial for most (and perhaps all) developmentally programmed somatic cell death
160
,
and another BH3‑only gene, ced‑13, which might have a role in stress‑induced cell
killing
161
. Zebrafish contain genes for nine versions of core BCL‑2 family members and
homologues of eight of the best‑studied BH3‑only genes
162
. Humans and mice have a
similar set of 12–13 structural homologues of BCL‑2 family proteins, which indicates
that family member organization is stable among mammals.
REVIEWS
NATURE REVIEWS
|
MOLECULAR CELL BIOLOGY VOLUME 9
|
JANUARY 2008
|
53
© 2008 Nature Publishing Group
Nature Reviews | Molecular Cell Biology
N terminus
C terminus
BAX
Soluble, cytosolic
conformation
in healthy cells
Model of membrane-inserted
apoptotic state of BAX
Conformational changes during apoptosis
Both pro- and anti-apoptotic BCL-2 family members
undergo dramatic conformational changes during apop-
tosis to unfold and insert deeply into the lipid bilayer (FIG. 5).
For example, BAX changes conformation to reveal a hidden
epitope in its N terminus
80,95,96
, shows increased proteolytic
sensitivity
81
and forms oligomeric complexes
97–99
around
the time it translocates to mitochondria.
Two steps in the activation process of BAX can be
discerned: an initial translocation to mitochondria, and
then the N-terminal conformational change that is likely
to be coupled to membrane insertion and oligomeriza-
tion. The translocational step appears to be reversible
under certain conditions
100
. Amino acids 14–23 in the
N terminus of BAX are hidden when the protein is
in its healthy cellconformation, but become exposed in
the early stages of apoptosis. The monoclonal antibody
6A7 binds to these amino acids and can be used to iden-
tify the rearrangement of this specific region of BAX
80
.
This BAX epitope has been crystallized bound to the
6A7 monoclonal antibody, revealing the large extent of
remodelling of the BAX N terminus that occurs dur-
ing apoptosis
101
. BAK also changes conformation
102
and
oligomerizes
99
during apoptosis. How many units of
BAX or BAK form these oligomers is still unclear, and it
appears that these oligomers are distributed over a wide
range of molecular weights.
BCL-2 also changes conformation during apoptosis
owing to binding of BH3-only proteins. Although
a model of this membrane-inserted form of BCL-2
(REF. 103) resembles that of BAX
104
, the anti-apoptotic
BCL-2 family proteins do not proceed to form oligomers,
unlike the pro-apoptotic family members. The conforma-
tion of anti-apoptotic BCL-2 family members in solution
(FIG. 3) probably represents the conformation found in
healthy cells, which does not bind to pro-apoptotic BAX.
However, BCL-2 and BCL-XL may bind BAX and BAK
after they insert into membranes, possibly when they
adopt conformations that resemble those induced by
detergents
85
, and may cap the BAX or BAK oligomers
and inhibit chain elongation
105
. One report concluded
that only the form of BCL-2 found in the early stages of
apoptosis could bind BAX and BAK and further inhibit
BAX or BAK oligomerization to promote cell survival
93
.
However, this would yield the counter-intuitive situation
in which BH3-only proteins could inhibit apoptosis by
promoting a change in the conformation of BCL-2 so
that it more actively inhibits BAX and BAK. Other work
found that the anti-apoptotic activity of BCL-W was
inactivated on insertion of BCL-W into the membrane
during apoptosis
92
.
Altering mitochondria
BCL-2 family members interact with mitochondria
either constitutively or on induction of apoptosis and,
although they might have activities in other cellular
compartments, it is clear that they regulate apoptosis by
their impact on the OMM.
Pro-apoptotic BCL-2 proteins induce cytochrome c
release. The OMM becomes permeable to soluble inter-
membrane space proteins at around the same time as
BAX is translocated and BAK undergoes conformational
change. Cytochrome c, DIABLO, adenylate kinase, the
Ser protease OMI, apoptosis-inducing factor (AIF),
deafness dystonia protein (DDP), endonuclease G and
a cleaved form of OPA1 (a mitochondrial dynamin-like
GTPase) have all been reported to be released from the
mitochondrial intermembrane space into the cytosol of
cells undergoing apoptosis
17,106–108
. Cytochrome c and
DIABLO release have been consistently shown to be
important for caspase activation.
Figure 5 | Conformational changes in BCL-2 family members during apoptosis. BAX undergoes extensive
conformational changes during the mitochondrial translocation process (see Supplementary information S1 (movie)). The
protein changes from a soluble cytoplasmic protein in healthy cells to one that appears to have at least three helices
inserted into the mitochondrial membrane in apoptotic cells. The C-terminal helix (blue) and the amphipathic helices (red)
in the soluble, cytoplasmic form of BAX (left) are thought to insert deeply into the mitochondrial membrane (right) during
early stages of apoptosis. In addition to changes in membrane topology, the N-terminal amino acids 13–27 (green) change
from a conformation that is buried in the protein folds and inaccessible to the monoclonal antibody 6A7 to a conformation
that is fully accessible to antibody binding during apoptosis. Immunostaining of cells with antibody 6A7 is a good marker
of the conformational change
95
. However, certain steps of this conformational change, such as the N terminus exposure,
may be reversible when apoptosis triggers are aborted. Other members of the BCL-2 family undergo similar changes in
conformation during apoptosis, including BAK, in which an N-terminal epitope is exposed. Interestingly, even anti-
apoptotic members of the BCL-2 family change conformation, and models of BCL-2 topology in the membrane that are
similar to those of BAX have been proposed
103
.
REVIEWS
54
|
JANUARY 2008
|
VOLUME 9 www.nature.com/reviews/molcellbio
© 2008 Nature Publishing Group
Erythropoiesis
The production of red blood
cells.
The initial finding that the structure of BCL-XL resem-
bles that of the translocation domain of diphtheria toxin
led to the proposal that BCL-XL might form pores in
membranes
29
. In trying to understand how pore-formation
activity is related to biological activity, a confusing factor
is that both pro- and anti-apoptotic BCL-2 family
members appear to be able to form membrane channels
in vitro
29,109,110
. Incubation of BAX with isolated mitochon-
dria induces cytochrome c release
111
, and incubating BAX
with liposomes allows the release of large (up to 10
6
Da)
dextran molecules
112
, which is consistent with the lip-
idic pores that were identified in lipid bilayer studies
113
.
BCL-XL inhibits BAX-induced cytochrome c release
from isolated mitochondria
111
and dextran release from
liposomes
70
. Furthermore, the BH3-only protein BID can
synergize with BAX to cause cytochrome c release in cell-
free assays, either by activating BAX or by preventing anti-
apoptotic BCL-2 family members from inhibiting BAX
and BAK. Thus, one model for BAX and BAK activation
is that they form large pores in the OMM that allow the
release of proteins into the cytosol, inducing caspase activ-
ation. However, the biochemical nature of this putative
BAX–BAK pore, such as the number of molecules of BAX
that comprise the pore, remains unknown.
Curiously, certain cell types (such as some neuro-
nal populations and cardiomyocytes) can survive the
cytochrome c release step, at least for a limited amount
of time
114,115
. In such cells, caspases might be stringently
regulated by caspase-inhibiting IAPs. In these cases,
apoptosis requires the release of DIABLO from mito-
chondria to relieve the IAP inhibition and thereby allow
caspase activation. This might be a specialized adapta-
tion of normally long-lived post-mitotic cells (which are
essential for animal survival) to supply these cells with
additional protection from cell-death activation and to
prevent their accidental death.
Roles in mitochondrial fragmentation and morphology.
Confocal and electron microscopy analyses of BAX
translocation to mitochondria reveal that the earliest
detectable form of ‘activated’ BAX does not localize to
the entire OMM, but is found concentrated at small
focal regions on the mitochondrial surface
116
. BAK also
moves into these ‘BAX foci’ during apoptosis induction.
This focal cluster form of BAX observed by microscopy
has an altered N-terminal conformation and probably
reflects the in situ state of BAX and BAK as oligomers.
These sites of BAX coalescence often develop into
mitochondrial division sites
117
, linking BAX to the pro-
motion of the mitochondrial fragmentation processes
that occur almost simultaneously with the release of
cytochrome c
118
(FIG. 2).
Inhibition of mitochondrial fission in vitro by
downregulation of the mitochondrial dynamin family
member DRP1 delays cytochrome c release and can
decrease caspase activation, which suggests that the
organelle division machinery somehow participates in
the regulation of apoptosis
119
. Deletion or mutation of
Drp1 in D. melanogaster
120,121
and drp-1 in C. elegans
122
also inhibits apoptosis in vivo (BOX 1). Conversely, inhi-
bition of mitochondrial fusion by loss of a different
mitochondrial dynamin family member, OPA1, induces
spontaneous apoptosis
123
.
Unexpectedly, healthy cells that lack both BAX and
BAK have altered mitochondrial morphology and slower
mitochondrial fusion rates, which indicates that BAX
and BAK affect mitochondrial morphogenesis machin-
eries even in the absence of apoptotic stimuli
124
. Recent
work showing that OPA1 controls mitochondrial cristae
formation and that tight cristae junctions can inhibit
cytochrome c release during apoptosis suggests how
mediators of mitochondrial fission and fusion might
have a role in cytochrome c release and apoptosis
125
. In
contrast to cytochrome c and OPA1, the release of other
mitochondrial intermembrane space proteins (such
as DIABLO) is not inhibited by DRP1 knockdown
126
,
which underscores the suggestion that the role of the
mitochondrial fission machinery in apoptosis might be
indirectly linked to the cytochrome c release step. Ectopic
expression of human BCL-XL and C. elegans ced-9 in
mammalian cells has been found to affect mitochondrial
morphogenesis, which shows that it is possible to sepa-
rate the process of organelle fusion regulation by BCL-2
family members from the regulation of apoptosis
127
.
Physiological roles of BCL-2 proteins
BCL-2 family members have essential roles in the mouse
from early embryogenesis through to adult tissue homeo-
stasis. The nervous system, haematopoietic tissues and
spermatogenesis are particularly dependent on BCL-2
family protein regulation (TABLE 1).
Anti-apoptotic BCL-2 proteins. MCL1 and BCL-XL
are both essential for normal embryogenesis. Mcl1
–/–
embryos die before implantation at the blastocyst stage
128
and Bcl-x
–/–
mice (in which the entire Bcl-x locus (incor-
porating Bcl-xl and Bcl-xs) was knocked out) survive only
until fetal day 13.5, displaying severe defects in erythropoi-
esis and neuronal development
129
. By contrast, although
BCL-2-deficient mice survive to birth, they have defects
in the immune system, hair follicles and renal epithe-
lial cells, and all succumb to polycystic kidney disease
by ~4–8 weeks of age (the age of death is influenced by
genetic background)
130,131
. Bcl-w-knockout male mice are
sterile owing to defective spermatogenesis, but otherwise
both females and males are developmentally normal
132
.
Analysis of the essential functions of A1 by gene targeting
is complicated by the fact that, in contrast to humans,
mice have four a1 genes. Mice that lack A1A are essen-
tially normal, but their granulocytes and mast cells
undergo apoptosis abnormally rapidly in culture
133,134
.
Pro-apoptotic BCL-2 proteins. Bax-knockout mice are
viable and females are fertile, but both males and females
have mild overgrowth of neurons and mild lymphoid
hyperplasia, and males have a severe defect in sperm-
cell differentiation, which results in sterility
135
. The
Bak
–/–
phenotype in mice is even less pronounced than
that of the Bax
–/–
mice: their fertility and most of their tis-
sues are normal
136
, although they do exhibit mild platelet
hypertrophy owing to a requirement for BAK to mediate
the turnover of these anuclear cellular fragments
137
.
REVIEWS
NATURE REVIEWS
|
MOLECULAR CELL BIOLOGY VOLUME 9
|
JANUARY 2008
|
55
© 2008 Nature Publishing Group
SLE-like autoimmune
disease
A rodent pathology that
resembles human systemic
lupus erythematosus, which is
commonly known as lupus.
Remarkably, however, Bax/Bak double knockout
mice display various severe defects, indicating extensive
redundancy in their activities
136
. A large fraction of Bax/
Bak double knockout mice die during embryogenesis
(particularly on an inbred C57BL/6 background; D.C.S.
Huang, unpublished observations) or perinatally (on a
mixed genetic background). The neonates display vari-
ous developmental deficits such as webs between their
digits, imperforate vaginas and abnormally increased
numbers of lymphoid and myeloid cells that are caused
by the persistent survival of cells that normally undergo
developmentally programmed death
136,138
. In vitro experi-
ments with cells from Bax/Bak double knockout mice
have shown that BAX and BAK are required for most
forms of stress-induced apoptosis
139
and that these cells
are even resistant to enforced expression of BH3-only
proteins
140,141
. These results demonstrate that BAX and
BAK are essential for apoptosis induction downstream
of the BH3-only proteins.
Importantly, the heart, liver, lungs and many other
organs develop normally in Bax/Bak double knockout
mice
136
. This might indicate that apoptosis, or at least
BAX/BAK-dependent apoptosis, is not crucial for normal
morphogenesis and normal cell turnover in these organs.
However, it remains possible that the closely related (but
only relatively poorly studied) protein BOK has a crucial
role in these tissues. It will therefore be informative to
generate Bok
–/–
, Bax/Bok and Bak/Bok double knockout
mice and, perhaps most importantly, Bax/Bak/Bok triple
knockout mice.
BH3-only proteins. Gene-targeting experiments have
also helped to define the essential functions of BH3-
only proteins (reviewed in REF. 142). Loss of BIM causes
abnormal accumulation of lymphoid and myeloid cells
and, on a mixed (C57BL/6x129SV) genetic background,
fatal SLE-like autoimmune disease
143
. BIM is crucial for
the deletion of autoreactive T and B cells
144,145
during
their development and for the termination of immune
responses
146
. In vitro experiments demonstrated that BIM
is essential for apoptosis that is induced by growth-factor
deprivation of a surprisingly broad range of cell types,
including lymphocytes
143
, osteoclasts
62
, mast cells
147
,
epithelial cells, endothelial cells
63
and neurons
148,149
.
BIM-deficient lymphocytes are also less vulnerable to
deregulated calcium flux and have only minor resistance
to γ-irradiation or treatment with glucocorticoids
143
.
Loss of BID has little effect on developmental apop-
tosis and, although it renders mice resistant to Fas-
induced hepatocyte apoptosis and fatal hepatitis
13,14
,
lymphoid cells from Bid
–/–
mice are normally sensitive
to Fas ligand
14
. PUMA, by contrast, is crucial for DNA-
damage-induced apoptosis, which is mediated by p53
(REFS 150–152). Curiously, although γ-radiation and UV
radiation both trigger apoptosis in a p53-dependent
manner, PUMA is essential for γ-radiation-induced
apoptosis and NOXA is essential for UV-radiation-
induced apoptosis within the same cell type (trans-
formed fibroblasts)
153
. This suggests that, depending on
the type of DNA damage and the nature of the molecular
mechanism of damage detection, p53 might be activated
in subtly different ways, thereby determining which of
its two pro-apoptotic BH3-only target genes is activated
preferentially. Alternatively, different forms of DNA
damage might activate distinct pathways that act in par-
allel with p53 signalling to determine whether PUMA,
NOXA or both are induced. PUMA is also crucial for
cell death that is induced by certain p53-independent
apoptotic stimuli, including cytokine deprivation or
treatment with glucocorticoids or phorbol ester
150,151
.
Mice that lack the BH3-only proteins that can only
bind some pro-survival proteins (BAD, BIK, HRK, BMF
or NOXA) have mild phenotypic abnormalities. This
is consistent with the hypothesis that these BH3-only
proteins are relatively weak killers compared to BIM,
PUMA or BID, which bind to anti-apoptotic BCL-2
family members more promiscuously. Mice that lack
BAD, BIK, HRK, BMF or NOXA are essentially normal
in appearance and are normally fertile. In BAD-deficient
mice, some cell types have subtle resistance to epidermal
growth factor or insulin growth factor deprivation
154
;
however, although these Bad
–/–
mice were reportedly
abnormally prone to lymphoma development, this could
not be reproduced in a subsequent study (P.N. Kelly
and A.S., unpublished observations). Neuronal popu-
lations from Hrk
–/–
mice exhibited some resistance to
nerve-growth-factor deprivation
155,156
, but this was less
pronounced than the protection afforded by loss of
BAX
157
, which indicates that other BH3-only proteins
are probably also involved.
Because many cells express more than one BH3-only
protein and several apoptotic stimuli can activate more
than one BH3-only protein, functional overlap appears
to be likely, and this has indeed been confirmed in early
studies on double knockout mice that lack two BH3-only
proteins. Although Bim
–/–
and Bik
–/–
male mice both have
normal spermatogenesis, severe defects that cause male
sterility became apparent in Bim/Bik double knockout
mice
158
. As in Bax
–/
males
135
, the failure to produce mature
sperm cells in Bim/Bik double knockout mice was a result
of the abnormal accumulation and persistence of imma-
ture progenitors, which prevent differentiating cells from
getting access to specialized niches on stromal cells.
Analysis of mice that lack both BIM and PUMA has
shown that these two BH3-only proteins are the most
crucial for apoptosis initiation in response to many death
stimuli in a broad range of cell types, particularly those
of haemopoietic origin
159
. For example, although loss of
either BIM or PUMA alone renders lymphoid and mye-
loid cells resistant to cytokine deprivation or treatment
with glucocorticoids, only the combined loss of both pro-
teins provided as much protection as the overexpression
of BCL-2 or the combined loss of BAX and BAK
159
.
BH3-only versus core BCL-2 proteins. The breeding of
mice that lack both a BH3-only protein and a BCL-2
pro-survival family member has helped to clarify func-
tional relationships between these proteins. Remarkably,
loss of a single allele of Bim prevents the fatal polycystic
kidney disease and lymphopoenia caused by loss of
BCL-2 (Bim
+/
Bcl-2
–/
mice), and loss of both Bim
alleles even prevents the abnormal death of melanocyte
REVIEWS
56
|
JANUARY 2008
|
VOLUME 9 www.nature.com/reviews/molcellbio
© 2008 Nature Publishing Group
progenitors and premature greying
131
. These results indi-
cate that when BCL-2 is absent in renal epithelial stem
cells, lymphoid cells and melanocyte progenitors, the
physiological levels of BIM are not sufficiently opposed
and cause abnormal apoptosis, presumably by neutralizing
the activity of other pro-survival BCL-2 family members,
such as BCL-XL or MCL1.
Concluding remarks
Our understanding of the regulation of BCL-2 family
members and their roles in tissue dynamics of mam-
mals has greatly expanded in recent years. BH3-only
proteins sense signals to induce apoptosis and relay this
information to core BCL-2 family members to initiate
cell death. BAX and BAK are induced to change confor-
mation and permeabilize the OMM. How BAX and BAK
function in this process remains unclear despite inten-
sive study. Difficulties in defining the structure of these
proteins after conformational change, oligomerization
and membrane insertion, as well as in determining their
intermolecular binding partners in membranes, has
impeded progress. The molecular trigger that induces
BAX translocation and BAK activation has so far also
eluded discovery.
The difference between anti- and pro-apoptotic BCL-2
family proteins needs to be defined both on a structural
and functional basis. One model to explain the difference is
that anti-apoptotic members can act as dominant-negative
inhibitors of the pro-apoptotic members. The functional
effect of BH3-only proteins binding to the anti-apoptotic
BCL-2 family members also deserves more study. Recent
advances in understanding the intermolecular interac-
tions among the family members, corroborated at the level
of animal studies, along with cell biology advances offer
abundant clues for deciphering the remaining mysteries
of cellular commitment to apoptosis.
1. Tsujimoto, Y., Cossman, J., Jaffe, E. & Croce, C. M.
Involvement of the bcl-2 gene in human follicular
lymphoma. Science 228, 1440–1443 (1985).
2. Bakhshi, A. et al. Cloning the chromosomal breakpoint
of t(14;18) human lymphomas: clustering around JH
on chromosome 14 and near a transcriptional unit on
18. Cell 41, 899–906 (1985).
3. Cleary, M. L., Smith, S. D. & Sklar, J. Cloning and
structural analysis of cDNAs for bcl-2 and a hybrid
bcl-2/immunoglobulin transcript resulting from the
t(14;18) translocation. Cell 47, 19–28 (1986).
References 1–3 describe the discovery of the
human BCL-2 gene.
4. Vaux, D. L., Cory, S. & Adams, J. M. Bcl-2 gene
promotes haemopoietic cell survival and cooperates
with c-myc to immortalize pre-B cells. Nature 335,
440–442 (1988).
Demonstrates that BCL-2 inhibits apoptotic cell
death, thereby identifying the first cell death
regulator, and shows that defects in apoptosis can
promote tumorigenesis.
5. Adams, J. M. & Cory, S. The Bcl-2 apoptotic switch in
cancer development and therapy. Oncogene 26,
1324–1337 (2007).
6. Evan, G. I. et al. Oncogene-dependent tumor
suppression: using the dark side of the force for cancer
therapy. Cold Spring Harb. Symp. Quant. Biol. 70,
263–273 (2005).
7. Strasser, A., Harris, A. W., Bath, M. L. & Cory, S.
Novel primitive lymphoid tumours induced in
transgenic mice by cooperation between myc and
bcl-2. Nature 348, 331–333 (1990).
8. Zha, H., Aime-Sempe, C., Sato, T. & Reed, J. C.
Proapoptotic protein Bax heterodimerizes with Bcl-2
and homodimerizes with Bax via a novel domain (BH3)
distinct from BH1 and BH2. J. Biol. Chem. 271,
7440–7444 (1996).
9. Aouacheria, A., Brunet, F. & Gouy, M. Phylogenomics
of life-or-death switches in multicellular animals: Bcl-2,
BH3-only, and BNip families of apoptotic regulators.
Mol. Biol. Evol. 22, 2395–2416 (2005).
10. Fesik, S. W. Promoting apoptosis as a strategy for
cancer drug discovery. Nature Rev. Cancer 5,
876–885 (2005).
11. Hakem, R. et al. Differential requirement for caspase 9
in apoptotic pathways in vivo. Cell 94, 339–352
(1998).
Demonstrated that caspase-9 is crucial for
apoptosis that is induced by intrinsic apoptotic
stimuli (such as growth-factor deprivation or DNA
damage) but is dispensable for death-receptor-
induced apoptosis.
12. Marsden, V. S. et al. Apoptosis initiated by
Bcl-2-regulated caspase activation independently of
the cytochrome c/Apaf-1/caspase-9 apoptosome.
Nature 419, 634–637 (2002).
13. Yin, X. M. et al. Bid-deficient mice are resistant to
Fas-induced hepatocellular apotosis. Nature 400,
886–891 (1999).
14. Kaufmann, T. et al. The BH3-only protein Bid is
dispensable for DNA damage- and replicative stress-
induced apoptosis or cell-cycle arrest. Cell 129,
423–433 (2007).
15. Willis, S. N. et al. Apoptosis initiated when BH3
ligands engage multiple Bcl-2 homologs, not Bax or
Bak. Science 315, 856–859 (2007).
16. Youle, R. J. Cell biology. Cellular demolition and the
rules of engagement. Science 315, 776–777
(2007).
17. Newmeyer, D. D. & Ferguson-Miller, S. Mitochondria:
releasing power for life and unleashing the
machineries of death. Cell 112, 481–490 (2003).
18. Chipuk, J. E., Bouchier-Hayes, L. & Green, D. R.
Mitochondrial outer membrane permeabilization
during apoptosis: the innocent bystander scenario.
Cell Death Differ. 13, 1396–1402 (2006).
19. Martinou, J. C. & Youle, R. J. Which came first, the
cytochrome c release or the mitochondrial fission?
Cell Death Differ. 13, 1291–1295 (2006).
20. Wang, X. The expanding role of mitochondria in
apoptosis. Genes Dev. 15, 2922–2933 (2001).
21. Shi, Y. Mechanical aspects of apoptosome assembly.
Curr. Opin. Cell Biol. 18, 677–684 (2006).
22. Hao, Z. et al. Specific ablation of the apoptotic
functions of cytochrome c reveals a differential
requirement for cytochrome c and Apaf-1 in apoptosis.
Cell 121, 579–591 (2005).
23. Okada, H. et al. Generation and characterization of
Smac/DIABLO-deficient mice. Mol. Cell. Biol. 22,
3509–3517 (2002).
24. Harlin, H., Reffey, S. B., Duckett, C. S., Lindsten, T. &
Thompson, C. B. Characterization of XIAP-deficient
mice. Mol. Cell. Biol. 21, 3604–3608 (2001).
25. Franchi, L., McDonald, C., Kanneganti, T. D., Amer, A.
& Nunez, G. Nucleotide-binding oligomerization
domain-like receptors: intracellular pattern recognition
molecules for pathogen detection and host defense.
J. Immunol. 177, 3507–3513 (2006).
26. Bruey, J. M. et al. Bcl-2 and Bcl-XL regulate
proinflammatory caspase-1 activation by interaction
with NALP1. Cell 129, 45–56 (2007).
27. Ekert, P. G. et al. Apaf-1 and caspase-9 accelerate
apoptosis, but do not determine whether factor-
deprived or drug-treated cells die. J. Cell Biol. 165,
835–842 (2004).
28. Marsden, V. S., Kaufmann, T., O’Reilly L, A., Adams,
J. M. & Strasser, A. Apaf-1 and caspase-9 are
required for cytokine withdrawal-induced apoptosis
of mast cells but dispensable for their functional
and clonogenic death. Blood 107, 1872–1877
(2006).
29. Muchmore, S. W. et al. X-ray and NMR structure of
human Bcl-xL, an inhibitor of programmed cell death.
Nature 381, 335 (1996).
Revealed the first 3D structure of a BCL-2 family
member.
30. Petros, A. M. et al. Solution structure of the
antiapoptotic protein bcl-2. Proc. Natl Acad. Sci. USA
98, 3012–3017 (2001).
31. Denisov, A. Y. et al. Solution structure of human
BCL-w: modulation of ligand binding by the C-terminal
helix. J. Biol. Chem. 278, 21124–21128 (2003).
32. Hinds, M. G. et al. The structure of Bcl-w reveals a
role for the C-terminal residues in modulating
biological activity. EMBO J. 22, 1497–1507
(2003).
33. Day, C. L. et al. Solution structure of prosurvival Mcl-1
and characterization of its binding by proapoptotic
BH3-only ligands. J. Biol. Chem. 280, 4738–4744
(2005).
34. Suzuki, M., Youle, R. J. & Tjandra, N. Structure of Bax:
co-regulation of dimer formation and intracellular
localization. Cell 103, 645–654 (2000).
This paper presents the 3D structure of BAX,
revealing that it is remarkably similar to that of
BCL-XL, although BAX promotes apoptosis
whereas BCL-XL promotes cell survival.
35. Moldoveanu, T. et al. The X-ray structure of a BAK
homodimer reveals an inhibitory zinc binding site.
Mol. Cell 24, 677–688 (2006).
36. McDonnell, J. M., Fushman, D., Milliman, C. L.,
Korsmeyer, S. J. & Cowburn, D. Solution structure of
the proapoptotic molecule BID: a structural basis for
apoptotic agonists and antagonists. Cell 96, 625–634
(1999).
37. Chou, J. J., Li, H., Salvesen, G. S., Yuan, J. &
Wagner, G. Solution structure of BID, an intracellular
amplifier of apoptotic signaling. Cell 96, 615–624
(1999).
38. Huang, Q., Petros, A. M., Virgin, H. W., Fesik, S. W. &
Olejniczak, E. T. Solution structure of a Bcl-2 homolog
from Kaposi sarcoma virus. Proc. Natl Acad. Sci. USA
99, 3428–3433 (2002).
39. Kvansakul, M. et al. A structural viral mimic of
prosurvival bcl-2: a pivotal role for sequestering
proapoptotic Bax and Bak. Mol. Cell 25, 933–942
(2007).
40. Douglas, A. E., Corbett, K. D., Berger, J. M.,
McFadden, G. & Handel, T. M. Structure of M11L: a
myxoma virus structural homolog of the apoptosis
inhibitor, Bcl-2. Protein Sci. 16, 695–703 (2007).
41. Aoyagi, M. et al. Vaccinia virus N1L protein resembles
a B cell lymphoma-2 (Bcl-2) family protein. Protein Sci.
16, 118–124 (2007).
42. Zha, J., Weiler, S., Oh, K. J., Wei, M. C. &
Korsmeyer, S. J. Posttranslational N-myristoylation of
BID as a molecular switch for targeting mitochondria
and apoptosis. Science 290, 1761–1765 (2000).
43. Sattler, M. et al. Structure of Bcl-xL–Bak peptide
complex: recognition between regulators of apoptosis.
Science 275, 983–986 (1997).
44. Petros, A. M. et al. Rationale for Bcl-xL/Bad peptide
complex formation from structure, mutagenesis, and
biophysical studies. Protein Sci. 9, 2528–2534
(2000).
45. Liu, X., Dai, S., Zhu, Y., Marrack, P. & Kappler, J. W.
The structure of a Bcl-xL/Bim fragment complex:
implications for Bim function. Immunity 19, 341–352
(2003).
46. Zhong, Q., Gao, W., Du, F. & Wang, X. Mule/ARF-BP1,
a BH3-only E3 ubiquitin ligase, catalyzes the
polyubiquitination of Mcl-1 and regulates apoptosis.
Cell 121, 1085–1095 (2005).
REVIEWS
NATURE REVIEWS
|
MOLECULAR CELL BIOLOGY VOLUME 9
|
JANUARY 2008
|
57
© 2008 Nature Publishing Group
47. Warr, M. R. et al. BH3-ligand regulates access of
MCL-1 to its E3 ligase. FEBS Lett. 579, 5603–5608
(2005).
48. Oberstein, A., Jeffrey, P. & Shi, Y. Crystal structure of
the BCL-XL–beclin 1 peptide complex: beclin 1 is a
novel BH3-only protein. J. Biol. Chem. 282,
13123–13132 (2007).
49. Hinds, M. G. et al. Bim, Bad and Bmf: intrinsically
unstructured BH3-only proteins that undergo a
localized conformational change upon binding to
prosurvival Bcl-2 targets. Cell Death Differ. 14,
128–136 (2007).
50. Grinberg, M. et al. tBID homooligomerizes in the
mitochondrial membrane to induce apoptosis. J. Biol.
Chem. 277, 12237–12245 (2002).
51. Schendel, S. L. et al. Ion channel activity of the BH3
only Bcl-2 family member, BID. J. Biol. Chem. 274,
21932–21936 (1999).
52. Wiens, M., Krasko, A., Muller, C. I. & Muller, W. E.
Molecular evolution of apoptotic pathways: cloning of
key domains from sponges (Bcl-2 homology domains
and death domains) and their phylogenetic
relationships. J. Mol. Evol. 50, 520–531 (2000).
53. Oda, E. et al. Noxa, a BH3-only member of the Bcl-2
family and candidate mediator of p53-induced
apoptosis. Science 288, 1053–1058 (2000).
54. Nakano, K. & Vousden, K. H. PUMA, a novel
proapoptotic gene, is induced by p53. Mol. Cell 7,
683–694 (2001).
55. Yu, J., Zhang, L., Hwang, P. M., Kinzler, K. W. &
Vogelstein, B. PUMA induces the rapid apoptosis of
colorectal cancer cells. Mol. Cell 7, 673–682 (2001).
56. Dijkers, P. F., Medema, R. H., Lammers, J. W.,
Koenderman, L. & Coffer, P. J. Expression of the pro-
apoptotic Bcl-2 family member Bim is regulated by the
forkhead transcription factor FKHR-L1. Curr. Biol. 10,
1201–1204 (2000).
57. Puthalakath, H. et al. ER stress triggers apoptosis by
activating BH3-only protein Bim via de-
phosphorylation and transcription induction. Cell 129,
1337–1349 (2007).
58. Zha, J., Harada, H., Yang, E., Jockel, J. &
Korsmeyer, S. J. 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
(1996).
59. Li, H., Zhu, H., Xu, C. J. & Yuan, J. Cleavage of BID
by caspase 8 mediates the mitochondrial damage in
the Fas pathway of apoptosis. Cell 94, 491–501
(1998).
60. Luo, X., Budihardjo, I., Zou, H., Slaughter, C. &
Wang, X. Bid, a Bcl2 interacting protein mediates
cytochrome c release from mitochondria in response
to activation of cell surface death receptors. Cell 94,
481–490 (1998).
61. Puthalakath, H., Huang, D. C., O’Reilly, L. A.,
King, S. M. & Strasser, A. The proapoptotic activity of
the Bcl-2 family member Bim is regulated by
interaction with the dynein motor complex. Mol. Cell
3, 287–296 (1999).
62. Akiyama, T. et al. Regulation of osteoclast apoptosis
by ubiquitylation of proapoptotic BH3-only Bcl-2
family member Bim. EMBO J. 22, 6653–6664
(2003).
63. Ley, R., Ewings, K. E., Hadfield, K. & Cook, S. J.
Regulatory phosphorylation of Bim: sorting out the
ERK from the JNK. Cell Death Differ. 12, 1008–1014
(2005).
64. Puthalakath, H. et al. Bmf: a proapoptotic BH3-only
protein regulated by interaction with the myosin V
actin motor complex, activated by anoikis. Science
293, 1829–1832 (2001).
65. Shimazu, T. et al. NBK/BIK antagonizes MCL-1 and
BCL-XL and activates BAK-mediated apoptosis in
response to protein synthesis inhibition. Genes Dev.
21, 929–941 (2007).
66. Grad, J. M., Zeng, X. R. & Boise, L. H. Regulation of
Bcl-xL: a little bit of this and a little bit of STAT. Curr.
Opin. Oncol. 12, 543–549 (2000).
67. Cuconati, A., Mukherjee, C., Perez, D. & White, E.
DNA damage response and MCL-1 destruction initiate
apoptosis in adenovirus-infected cells. Genes Dev. 17,
2922–2932 (2003).
68. Letai, A. et al. Distinct BH3 domains either sensitize
or activate mitochondrial apoptosis, serving as
prototype cancer therapeutics. Cancer Cell 2,
183–192 (2002).
69. Chen, L. et al. Differential targeting of prosurvival
Bcl-2 proteins by their BH3-only ligands allows
complementary apoptotic function. Mol. Cell 17,
393–403 (2005).
70. Kuwana, T. et al. BH3 domains of BH3-only proteins
differentially regulate Bax-mediated mitochondrial
membrane permeabilization both directly and
indirectly. Mol. Cell 17, 525–535 (2005).
71. Kim, H. et al. Hierarchical regulation of mitochondrion-
dependent apoptosis by BCL-2 subfamilies. Nature
Cell Biol. 8, 1348–1358 (2006).
72. Certo, M. et al. Mitochondria primed by death signals
determine cellular addiction to antiapoptotic BCL-2
family members. Cancer Cell 9, 351–365 (2006).
73. Willis, S. N. et al. Proapoptotic Bak is sequestered by
Mcl-1 and Bcl-xL, but not Bcl-2, until displaced by
BH3-only proteins. Genes Dev. 19, 1294–1305
(2005).
74. Walensky, L. D. et al. A stapled BID BH3 helix directly
binds and activates BAX. Mol. Cell 24, 199–210
(2006).
75. Nguyen, M., Millar, D. G., Yong, V. W., Korsmeyer, S. J.
& Shore, G. C. Targeting of Bcl-2 to the mitochondrial
outer membrane by a COOH-terminal signal anchor
sequence. J. Biol. Chem. 268, 25265–25268 (1993).
76. Lithgow, T., van Driel, R., Bertram, J. F. & Strasser, A.
The protein product of the oncogene bcl-2 is a
component of the nuclear envelope, the endoplasmic
reticulum, and the outer mitochondrial membrane.
Cell Growth Differ. 5, 411–417 (1994).
77. Heath-Engel, H. M. & Shore, G. C. Regulated targeting
of Bax and Bak to intracellular membranes during
apoptosis. Cell Death Differ. 13, 1277–1280 (2006).
78. Pinton, P. & Rizzuto, R. Bcl-2 and Ca
2+
homeostasis in
the endoplasmic reticulum. Cell Death Differ. 13,
1409–1418 (2006).
79. Hsu, Y.-T., Wolter, K. & Youle, R. J. Cytosol to
membrane redistribution of members of the Bcl-2
family during apoptosis. Proc. Natl Acad. Sci. USA 94,
3668–3672 (1997).
80. Hsu, Y. T. & Youle, R. J. Bax in murine thymus is a
soluble monomeric protein that displays differential
detergent-induced conformations. J. Biol. Chem. 273,
10777–10783 (1998).
81. Goping, I. S. et al. Regulated targeting of BAX to
mitochondria. J. Cell Biol. 143, 207–215 (1998).
82. Wolter, K. G. et al. Movement of Bax from the cytosol
to mitochondria. J. Cell Biol. 139, 1281–1292 (1997).
83. Cartron, P. F. et al. Involvement of the N-terminus of
Bax in its intracellular localization and function.
FEBS Lett. 512, 95–100 (2002).
84. Gao, S., Fu, W., Durrenberger, M., De Geyter, C. &
Zhang, H. Membrane translocation and
oligomerization of hBok are triggered in response to
apoptotic stimuli and Bnip3. Cell. Mol. Life Sci. 62,
1015–1024 (2005).
85. Hsu, Y.-T. & Youle, R. J. Nonionic detergent induced
dimerization of members of the Bcl-2 family. J. Biol.
Chem. 272, 13829–13834 (1997).
86. Cheng, E. H., Sheiko, T. V., Fisher, J. K., Craigen, W. J.
& Korsmeyer, S. J. VDAC2 inhibits BAK activation and
mitochondrial apoptosis. Science 301, 513–517
(2003).
87. Setoguchi, K., Otera, H. & Mihara, K. Cytosolic factor-
and TOM-independent import of C-tail-anchored
mitochondrial outer membrane proteins. EMBO J. 25,
5635–5647 (2006).
88. Baines, C. P., Kaiser, R. A., Sheiko, T., Craigen, W. J. &
Molkentin, J. D. Voltage-dependent anion channels
are dispensable for mitochondrial-dependent cell
death. Nature Cell Biol. 9, 550–555 (2007).
89. Jeong, S. Y. et al. Bcl-x(L) sequesters its C-terminal
membrane anchor in soluble, cytosolic homodimers.
EMBO J. 23, 2146–2155 (2004).
90. Nijhawan, D. et al. Elimination of Mcl-1 is required for
the initiation of apoptosis following ultraviolet
irradiation. Genes Dev. 17, 1475–1486 (2003).
91. Hausmann, G. et al. Pro-apoptotic apoptosis protease-
activating factor 1 (Apaf-1) has a cytoplasmic
localization distinct from Bcl-2 or Bcl-x(L). J. Cell Biol.
149, 623–634 (2000).
92. Wilson-Annan, J. et al. Proapoptotic BH3-only
proteins trigger membrane integration of prosurvival
Bcl-w and neutralize its activity. J. Cell Biol. 162,
877–887 (2003).
93. Kim, P. K., Annis, M. G., Dlugosz, P. J., Leber, B. &
Andrews, D. W. During apoptosis Bcl-2 changes
membrane topology at both the endoplasmic reticulum
and mitochondria. Mol. Cell 14, 523–529 (2004).
94. Strasser, A., O’Connor, L. & Dixit, V. M. Apoptosis
signaling. Annu. Rev. Biochem. 69, 217–245 (2000).
95. Nechushtan, A., Smith, C. L., Hsu, Y.-T. & Youle, R. J.
Conformation of the Bax C-terminus regulates
subcellular location and cell death. EMBO J. 18,
2330–2341 (1999).
96. Desagher, S. et al. Bid-induced conformational change
of Bax is responsible for mitochondrial cytochrome c
release during apoptosis. J. Cell Biol. 144, 891–901
(1999).
97. Tan, Y. J., Beerheide, W. & Ting, A. E. Biophysical
characterization of the oligomeric state of Bax and its
complex formation with Bcl-XL. Biochem. Biophys.
Res. Commun. 255, 334–339 (1999).
98. Antonsson, B., Montessuit, S., Lauper, S., Eskes, R. &
Martinou, J. C. Bax oligomerization is required for
channel-forming activity in liposomes and to trigger
cytochrome c release from mitochondria. Biochem. J.
345, 271–278 (2000).
99. Mikhailov, V. et al. Association of Bax and Bak homo-
oligomers in mitochondria. Bax requirement for Bak
reorganization and cytochrome c release. J. Biol.
Chem. 278, 5367–5376 (2003).
100. Valentijn, A. J., Metcalfe, A. D., Kott, J., Streuli, C. H.
& Gilmore, A. P. Spatial and temporal changes in Bax
subcellular localization during anoikis. J. Cell Biol.
162, 599–612 (2003).
101. Peyerl, F. W. et al. Elucidation of some Bax
conformational changes through crystallization of an
antibody-peptide complex. Cell Death Differ. 14,
447–452 (2006).
102. Griffiths, G. J. et al. Cell damage-induced
conformational changes of the pro-apoptotic protein
Bak in vivo precede the onset of apoptosis. J. Cell Biol.
144, 903–914 (1999).
103. Dlugosz, P. J. et al. Bcl-2 changes conformation to
inhibit Bax oligomerization. EMBO J. 25, 2287–2296
(2006).
104. Annis, M. G. et al. Bax forms multispanning
monomers that oligomerize to permeabilize
membranes during apoptosis. EMBO J. 24,
2096–2103 (2005).
105. Ruffolo, S. C. & Shore, G. C. BCL-2 selectively interacts
with the BID-induced open conformer of BAK,
inhibiting BAK auto-oligomerization. J. Biol. Chem.
278, 25039–25045 (2003).
106. Ekert, P. G. & Vaux, D. L. The mitochondrial death
squad: hardened killers or innocent bystanders? Curr.
Opin. Cell Biol. 17, 626–630 (2005).
107. Green, D. R. & Kroemer, G. The pathophysiology of
mitochondrial cell death. Science 305, 626–629
(2004).
108. Arnoult, D., Grodet, A., Lee, Y. J., Estaquier, J. &
Blackstone, C. Release of OPA1 during apoptosis
participates in the rapid and complete release of
cytochrome c and subsequent mitochondrial
fragmentation. J. Biol. Chem. 280, 35742–35750
(2005).
109. Antonsson, B. et al. Inhibition of Bax channel-
forming activity by Bcl-2. Science 277, 370–372
(1997).
110. Minn, A. J. et al. Bcl-xL forms an ion channel in
synthetic lipid membranes. Nature 385, 353357
(1997).
111. Jurgensmeier, J. M. et al. Bax directly induces release
of cytochrome c from isolated mitochondria. Proc. Natl
Acad. Sci. USA 95, 4997–5002 (1998).
112. Kuwana, T. et al. Bid, Bax, and lipids cooperate to
form supramolecular openings in the outer
mitochondral membreane. Cell 111, 331–342
(2002).
113. Basanez, G. et al. Full length Bax disrupts planar
phospholipid membranes. Proc. Natl Acad. Sci. USA
96, 5492–5497 (1999).
114. Martinou, I. et al. The release of cytochrome c from
mitochondria during apoptosis of NGF-deprived
sympathetic neurons is a reversible event. J. Cell Biol.
144, 883–889 (1999).
115. Potts, M. B., Vaughn, A. E., McDonough, H.,
Patterson, C. & Deshmukh, M. Reduced Apaf-1 levels
in cardiomyocytes engage strict regulation of
apoptosis by endogenous XIAP. J. Cell Biol. 171,
925–930 (2005).
116. Nechushtan, A., Smith, C. L., I., L., Yoon, S. H. &
Youle, R. J. Bax and Bak coalesce into novel
mitochondria-associated clusters during apoptosis.
J. Cell Biol. 153, 1265–1276 (2001).
117. Karbowski, M. et al. Spatial and temporal association
of Bax with mitochondrial fission sites, Drp1, and
Mfn2 during apoptosis. J. Cell Biol. 159, 931–938
(2002).
118. Youle, R. J. & Karbowski, M. Mitochondrial fission in
apoptosis. Nature Rev. Mol. Cell Biol. 6, 657–663
(2005).
119. Frank, S. et al. The role of dynamin-related protein 1,
a mediator of mitochondrial fission, in apoptosis.
Dev. Cell 1, 515–525 (2001).
REVIEWS
58
|
JANUARY 2008
|
VOLUME 9 www.nature.com/reviews/molcellbio
© 2008 Nature Publishing Group
120. Goyal, G., Fell, B., Sarin, A., Youle, R. J. & Sriram, V.
Role of mitochondrial remodeling in programmed cell
death in Drosophila melanogaster. Dev. Cell 12,
807–816 (2007).
121. Abdelwahid, E. et al. Mitochondrial disruption in
Drosophila apoptosis. Dev. Cell 12, 793–806
(2007).
122. Jagasia, R., Grote, P., Westermann, B. & Conradt, B.
DRP-1-mediated mitochondrial fragmentation during
EGL-1-induced cell death in C. elegans. Nature 433,
754–760 (2005).
123. Olichon, A. et al. Loss of OPA1 perturbates the
mitochondrial inner membrane structure and integrity,
leading to cytochrome c release and apoptosis. J. Biol.
Chem. 278, 7743–7746 (2003).
124. Karbowski, M., Norris, K. L., Cleland, M. M., Jeong,
S. Y. & Youle, R. J. Role of Bax and Bak in
mitochondrial morphogenesis. Nature 443, 658–662
(2006).
125. Cipolat, S. et al. Mitochondrial rhomboid PARL
regulates cytochrome c release during apoptosis via
OPA1-dependent cristae remodeling. Cell 126,
163–175 (2006).
126. Parone, P. A. et al. Inhibiting the mitochondrial
fission machinery does not prevent Bax/Bak-
dependent apoptosis. Mol. Cell. Biol. 26,
7397–7408 (2006).
127. Delivani, P., Adrain, C., Taylor, R. C., Duriez, P. J. &
Martin, S. J. Role for CED-9 and Egl-1 as regulators of
mitochondrial fission and fusion dynamics. Mol. Cell
21, 761–773 (2006).
128. Rinkenberger, J. L., Horning, S., Klocke, B., Roth, K. &
Korsmeyer, S. J. Mcl-1 deficiency results in peri-
implantation embryonic lethality. Genes Dev. 14,
23–27 (2000).
Shows that the anti-apoptotic BCL-2 family
member MCL1 is required for early steps in mouse
embryonic development.
129. Motoyama, N. et al. Massive cell death of immature
hematopoietic cells and neurons in Bcl-x-deficient
mice. Science 267, 1506–1510 (1995).
Shows that BCL-XL is essential for the survival of
immature erythroid progenitors and neuronal cells
during mouse embryonic development.
130. Veis, D. J., Sorenson, C. M., Shutter, J. R. &
Korsmeyer, S. J. Bcl-2-deficient mice demonstrate
fulminant lymphoid apoptosis, polycystic kidneys, and
hypopigmented hair. Cell 75, 229–240 (1993).
Shows that BCL-2 is essential for survival of renal
epithelial progenitors, mature lymphocytes and
melanocyte progenitors in the mouse.
131. Bouillet, P., Cory, S., Zhang, L. C., Strasser, A. &
Adams, J. M. Degenerative disorders caused by Bcl-2
deficiency prevented by loss of its BH3-only antagonist
Bim. Dev. Cell 1, 645–653 (2001).
132. Print, C. G. et al. Apoptosis regulator Bcl-w is essential
for spermatogenesis but appears otherwise
redundant. Proc. Natl Acad. Sci. USA 95,
12424–12431 (1998).
133. Hamasaki, A. et al. Accelerated neutrophil apoptosis
in mice lacking A1-a, a subtype of the Bcl-2-related
A1 gene. J. Exp. Med. 188, 1985–1992 (1998).
134. Xiang, Z. et al. Essential role of the prosurvival Bcl-2
homologue A1 in mast cell survival after allergic
activation. J. Exp. Med. 194, 1561–1569
(2001).
135. Knudson, C. M., Tung, K. S., Tourtellotte, W. G.,
Brown, G. A. & Korsmeyer, S. J. Bax-deficient mice
with lymphoid hyperplasia and male germ cell death.
Science 270, 96–99 (1995).
136. Lindsten, T. et al. The combined functions of
proapoptotic Bcl-2 family members Bak and Bax are
essential for normal development of multiple tissues.
Mol. Cell 6, 1389–1399 (2000).
137. Mason, K. D. et al. Programmed anuclear cell death
delimits platelet life span. Cell 128, 1173–1186
(2007).
138. Rathmell, J. C., Lindsten, T., Zong, W. X., Cinalli, R. M.
& Thompson, C. B. Deficiency in Bak and Bax perturbs
thymic selection and lymphoid homeostasis. Nature
Immunol. 3, 932–939 (2002).
Along with reference 136, demonstrates that BAX
and BAK have largely overlapping functions in
developmentally programmed cell death and
stress-induced apoptosis.
139. Wei, M. C. et al. Proapoptotic BAX and BAK: a
requisite gateway to mitochondrial dysfunction and
death. Science 292, 727–730 (2001).
140. Zong, W. X., Lindsten, T., Ross, A. J., MacGregor, G. R.
& Thompson, C. B. BH3-only proteins that bind pro-
survival Bcl-2 family members fail to induce apoptosis
in the absence of Bax and Bak. Genes Dev. 15,
1481–1486 (2001).
141. Cheng, E. H. et al. BCL-2, BCL-X(L) sequester BH3
domain-only molecules preventing BAX- and BAK-
mediated mitochondrial apoptosis. Mol. Cell 8,
705–711 (2001).
References 140 and 141 demonstrate that BAX
and/or BAK are required for apoptosis induced by
BH3-only proteins.
142. Strasser, A. The role of BH3-only proteins in the
immune system. Nature Rev. Immunol. 5, 189–200
(2005).
143. Bouillet, P. et al. Proapoptotic Bcl-2 relative Bim
required for certain apoptotic responses, leukocyte
homeostasis, and to preclude autoimmunity. Science
286, 1735–1738 (1999).
Provides the first evidence that a BH3-only protein,
BIM, is essential for developmentally programmed
cell death in mammals.
144. Bouillet, P. et al. BH3-only Bcl-2 family member Bim is
required for apoptosis of autoreactive thymocytes.
Nature 415, 922–926 (2002).
145. Enders, A. et al. Loss of the pro-apoptotic BH3-only
Bcl-2 family member Bim inhibits BCR stimulation-
induced apoptosis and deletion of autoreactive B cells.
J. Exp. Med. 198, 1119–1126 (2003).
146. Pellegrini, M., Belz, G., Bouillet, P. & Strasser, A.
Shutdown of an acute T cell immune response to viral
infection is mediated by the proapoptotic Bcl-2
homology 3-only protein Bim. Proc. Natl Acad. Sci.
USA 100, 14175–14180 (2003).
147. Alfredsson, J., Puthalakath, H., Martin, H., Strasser, A.
& Nilsson, G. Proapoptotic Bcl-2 family member Bim
is involved in the control of mast cell survival and is
induced together with Bcl-XL upon IgE-receptor
activation. Cell Death Differ. 12, 136–144
(2005).
148. Putcha, G. V. et al. JNK-mediated BIM
phosphorylation potentiates BAX-dependent
apoptosis. Neuron 38, 899–914 (2003).
149. Whitfield, J., Neame, S. J., Paquet, L., Bernard, O. &
Ham, J. Dominant-negative c-Jun promotes neuronal
survival by reducing BIM expression and inhibiting
mitochondrial cytochrome c release. Neuron 29,
629–643 (2001).
150. Villunger, A. et al. p53- and drug-induced apoptotic
responses mediated by BH3-only proteins Puma and
Noxa. Science 302, 1036–1038 (2003).
151. Jeffers, J. R. et al. Puma is an essential mediator of
p53-dependent and -independent apoptotic pathways.
Cancer Cell 4, 321–328 (2003).
Together with reference 150, shows that the BH3-
only protein PUMA is essential for p53-mediated
apoptosis triggered by DNA damage and also
for apoptosis that is induced by certain p53-
independent stimuli, such as cytokine deprivation
or treatment with glucocorticoids.
152. Erlacher, M. et al. BH3-only proteins Puma and Bim
are rate-limiting for γ-radiation- and glucocorticoid-
induced apoptosis of lymphoid cells in vivo. Blood
106, 4131–4138 (2005).
153. Naik, E., Michalak, E. M., Villunger, A., Adams, J. M.
& Strasser, A. Ultraviolet radiation triggers apoptosis
of fibroblasts and skin keratinocytes mainly via the
BH3-only protein Noxa. J. Cell Biol. 176, 415–424
(2007).
154. Ranger, A. M. et al. Bad-deficient mice develop diffuse
large B cell lymphoma. Proc. Natl Acad. Sci. USA 100,
9324–9329 (2003).
155. Imaizumi, K. et al. Critical role for DP5/Harakiri, a
Bcl-2 homology domain 3-only Bcl-2 family member,
in axotomy-induced neuronal cell death. J. Neurosci.
24, 3721–3725 (2004).
156. Coultas, L. et al. Pro-apoptotic BH3-only Bcl-2 family
member Hrk/DP5 contributes to the apoptosis of
select neuronal populations but is dispensible for
hemopoetic cell apoptosis. J. Cell Sci. 120,
2044–2052 (2007).
157. Deckwerth, T. L. et al. BAX is required for neuronal
death after trophic factor deprivation and during
development. Neuron 17, 401–411 (1996).
158. Coultas, L. et al. Concomitant loss of proapoptotic
BH3-only Bcl-2 antagonists Bik and Bim arrests
spermatogenesis. EMBO J. 24, 3963–3973 (2005).
159. Erlacher, M. et al. Puma cooperates with Bim, the
rate-limiting BH3-only protein in cell death during
lymphocyte development, in apoptosis induction.
J. Exp. Med. 203, 2939–2951 (2006).
160. Conradt, B. & Horvitz, H. R. The C. elegans protein
EGL-1 is required for programmed cell death and
interacts with the Bcl-2-like protein CED-9. Cell 93,
519–529 (1998).
Shows that the BH3-only protein EGL-1 is essential
for developmentally programmed cell death in
C. elegans.
161. Schumacher, B. et al. C. elegans ced-13 can promote
apoptosis and is induced in response to DNA damage.
Cell Death Differ. 12, 153–161 (2005).
162. Kratz, E. et al. Functional characterization of the Bcl-2
gene family in the zebrafish. Cell Death Differ. 13,
1631–1640 (2006).
163. Hengartner, M. O. & Horvitz, H. R. Activation of
C. elegans cell death protein CED-9 by an amino-acid
substitution in a domain conserved in Bcl-2. Nature
369, 318–320 (1994).
Shows that CED-9, which is essential for cell
survival during development in C. elegans, is a
homologue of mammalian BCL-2, indicating that
the control of apoptosis is evolutionarily conserved.
164. Cheng, E. H. et al. Conversion of Bcl-2 to a Bax-like
death effector by caspases. Science 278, 1966–1968
(1997).
165. Lin, B. et al. Conversion of Bcl-2 from protector to
killer by interaction with nuclear orphan receptor
Nur77/TR3. Cell 116, 527–540 (2004).
166. Sevrioukov, E. A. et al. Drosophila Bcl-2 proteins
participate in stress-induced apoptosis, but are not
required for normal development. Genesis 45,
184–193 (2007).
DATABASES
Entrez Gene: http://www.ncbi.nlm.nih.gov/entrez/query.
fcgi?db=gene
BCL-2
UniProtKB: http://beta.uniprot.org/uniprot
APAF1 | BAD | BAK | BAX | A1 | BCL-B | BCL-W | BCL-XL | BID |
BIK | BIM | BMF | BOK | caspase-8 | caspase-9 | DIABLO | HRK |
MCL1 | PUMA | XIAP
FURTHER INFORMATION
Richard J. Youle’s homepage:
http://neuroscience.nih.gov/Lab.asp?Org_ID=81
Andreas Strasser’s homepage:
http://www.wehi.edu.au/facweb/indexresearch.php?id=24
SUPPLEMENTARY INFORMATION
See online article: S1 (movie) | S2 (table)
ALL LINKS ARE ACTIVE IN THE ONLINE PDF
REVIEWS
NATURE REVIEWS
|
MOLECULAR CELL BIOLOGY VOLUME 9
|
JANUARY 2008
|
59
© 2008 Nature Publishing Group
... Within mammals, B-cell lymphoma-2 (Bcl-2) family proteins such as Bcl-2, Bcl-xL, Bcl-w, Mcl-1, and A1 play a role in apoptosis. With the exception of Bim, Bid, and Bad, which are BH3only proteins, Bcl-2 proteins consist of four homologous domains (BH1, BH2, BH3, and BH4) [69][70][71][72]. Alterations in genetics associated with cancer and the development of tumors frequently impact the control of cell death to stimulate cell proliferation. ...
Article
Full-text available
Despite decades of basic and clinical research and trials of promising new therapies, cancer remains a major cause of morbidity and mortality due to the emergence of drug resistance to anticancer drugs. These resistance events have a very well-understood underlying mechanism, and their therapeutic relevance has long been recognized. Thus, drug resistance continues to be a major obstacle to providing cancer patients with the intended “cure”. PAQR4 (Progestin and AdipoQ Receptor Family Member 4) gene is a recently identified novel protein-coding gene associated with various human cancers and acts through different signaling pathways. PAQR4 has a significant influence on multiple proteins that may regulate various gene expressions and may develop chemoresistance. This review discusses the roles of PAQR4 in tumor immunity, carcinogenesis, and chemoresistance. This paper is the first review, discussing PAQR4 in the pathogenesis of cancer. The review further explores the PAQR4 as a potential target in various malignancies.
... .‫ا:5ا‬Members‫ا‬of‫ا‬BcI-2‫ا‬famiIy‫ا‬of‫ا‬proteins ‫ا‬[32] ...
... There are three phases of apoptosis: induction, effect, and degradation. The Bcl-2 family is mainly involved in the effect phase [56], including both anti-apoptotic proteins (e.g., Bcl-2, Bcl-W, Bcl-XL, and Mcl-1) and pro-apoptotic proteins (e.g., Bad, Bax, Bcl-Xs, Bak, and Bid), which jointly regulate the process of apoptosis [57][58][59]. During IR, the Bad protein accumulates in the cytoplasm and then combines with Bcl-2 and Bcl-XL to process Bax and Bak, inducing the release of downstream pro-apoptotic proteins [60] (Fig. 5). ...
... It was reported that many genes were involved in the tumorigenesis. Bcl-2 is a vital gene that suppresses pro-apoptotic signals to prevent cell apoptosis, and its excessive expression will cause out of control cell proliferation [53]. As the most easily mutated proto-oncogene, the Kras gene mutation always participates in adenocarcinoma progression [54]. ...
Article
Full-text available
Chemotherapy exhibits numerous side effects in anti-tumour therapy. The clinical experiments indicated that deuterium-depleted water (DDW) monotherapy or in combination with chemotherapy was beneficial in inhibiting cancer development. To further understand the potential mechanism of DDW in cancer therapy, we performed a systematic review. The data from experiments published over the past 15 years were included. PubMed, Cochrane and Web of Science (January 2008 to November 2023) were systemically searched. Fifteen studies qualified for review, including fourteen in vivo and in vitro trials and one interventional trial. The results showed that DDW alone or in combination with chemotherapy effectively inhibited cancer progression in most experiments. The combination treatment enhances the therapeutic effect on cancer compared with chemotherapeutic monotherapy. The inhibitory role of DDW in tumours is through regulating the reactive oxygen species (ROS)-related genes in Kelch-like ECH-associated protein 1 (Keap 1) and Nuclear erythroid 2-related factor 2 (Nrf2) signalling pathways, further controlling ROS production. An abnormal amount of ROS can inhibit the tumour progression. More extensive randomized controlled trials should be conducted to evaluate the accurate effect of DDW in Keap1-Nrf2 signalling pathways.
... In addition, we observed a decrease in the expression of the prosurvival mediator BcL2, which belongs to a family of mediators involved in regulating cell survival 74 . Although this reduction was less pronounced https://doi.org/10.1038/s42003-024-05922-2 ...
Article
Full-text available
COVID-19, caused by SARS-CoV-2, can lead to a severe inflammatory disease characterized by significant lymphopenia. However, the underlying cause for the depletion of T-cells in COVID-19 patients remains incompletely understood. In this study, we assessed the presence of different T-cell subsets in the progression of COVID-19 from mild to severe disease, with a focus on TCF1 expressing progenitor T-cells that are needed to replenish peripheral T-cells during infection. Our results showed a preferential decline in TCF1+ progenitor CD4 and CD8+ T-cells with disease severity. This decline was seen in various TCF1+ subsets including naive, memory and effector-memory cells, and surprisingly, was accompanied by a loss in cell division as seen by a marked decline in Ki67 expression. In addition, TCF1+ T-cells showed a reduction in pro-survival regulator, BcL2, and the appearance of a new population of TCF1 negative caspase-3 expressing cells in peripheral blood from patients with severe disease. The decline in TCF1+ T-cells was also seen in a subgroup of severe patients with vitamin D deficiency. Lastly, we found that sera from severe patients inhibited TCF1 transcription ex vivo which was attenuated by a blocking antibody against the cytokine, interleukin-12 (IL12). Collectively, our findings underscore the potential significance of TCF1+ progenitor T-cells in accounting for the loss of immunity in severe COVID-19 and outline an array of markers that could be used to identify disease progression.
Article
Full-text available
Cellular demise is a pivotal event in both developmental processes and disease states, with mitochondrial regulation playing an essential role. Traditionally, cell death was categorized into distinct types, considered to be linear and mutually exclusive pathways. However, the current understanding has evolved to recognize the complex and interconnected mechanisms of cell death, especially within apoptosis, pyroptosis, and necroptosis. Apoptosis, pyroptosis, and necroptosis are governed by intricate molecular pathways, with mitochondria acting as central decision-makers in steering cells towards either apoptosis or pyroptosis through various mediators. The choice between apoptosis and necroptosis is often determined by mitochondrial signaling and is orchestrated by specific proteins. The molecular dialogue and the regulatory influence of mitochondria within these cell death pathways are critical research areas. Comprehending the shared elements and the interplay between these death modalities is crucial for unraveling the complexities of cellular demise.
Article
Full-text available
Anticancer immune surveillance and immunotherapies trigger activation of cytotoxic cytokine signaling, including tumor necrosis factor-α (TNF-α) and TNF-related apoptosis-inducing ligand (TRAIL) pathways. The pro-inflammatory cytokine TNF-α may be secreted by stromal cells, tumor-associated macrophages, and by cancer cells, indicating a prominent role in the tumor microenvironment (TME). However, tumors manage to adapt, escape immune surveillance, and ultimately develop resistance to the cytotoxic effects of TNF-α. The mechanisms by which cancer cells evade host immunity is a central topic of current cancer research. Resistance to TNF-α is mediated by diverse molecular mechanisms, such as mutation or downregulation of TNF/TRAIL receptors, as well as activation of anti-apoptotic enzymes and transcription factors. TNF-α signaling is also mediated by sphingosine kinases (SphK1 and SphK2), which are responsible for synthesis of the growth-stimulating phospholipid, sphingosine-1-phosphate (S1P). Multiple studies have demonstrated the crucial role of S1P and its transmembrane receptors (S1PR) in both the regulation of inflammatory responses and progression of cancer. Considering that the SphK/S1P/S1PR axis mediates cancer resistance, this sphingolipid signaling pathway is of mechanistic significance when considering immunotherapy-resistant malignancies. However, the exact mechanism by which sphingolipids contribute to the evasion of immune surveillance and abrogation of TNF-α-induced apoptosis remains largely unclear. This study reviews mechanisms of TNF-α-resistance in cancer cells, with emphasis on the pro-survival and immunomodulatory effects of sphingolipids. Inhibition of SphK/S1P-linked pro-survival branch may facilitate reactivation of the pro-apoptotic TNF superfamily effects, although the role of SphK/S1P inhibitors in the regulation of the TME and lymphocyte trafficking should be thoroughly assessed in future studies.
Article
Full-text available
Most members of the Bcl-2 protein family of apoptosis regulating proteins contain two evolutionarily conserved domains, termed BH1 and BH2. Both BH1 and BH2 in the Bcl-2 protein are required for its function as an inhibitor of cell death and for heterodimerization with the proapoptotic protein Bax. In this report, we mapped the region in Bax required for heterodimerization with Bcl-2 and homodimerization with Bax, using yeast two-hybrid and in vitro protein-protein interaction assays. Neither the BH1 nor the BH2 domain of Bax was required for binding to the wild-type Bcl-2 and Bax proteins. Moreover, Bax (ΔBH1) and Bax (ΔBH2) mutant proteins bound efficiently to themselves and each other, further confirming the lack of requirement for BH1 and BH2 for Bax/Bax homodimerization. Bax/Bax homodimerization was not dependent on the inclusion of the NH-terminal 58 amino acids of the Bax protein in each dimerization partner, unlike Bcl-2/Bcl-2 homodimers which involve head-to-tail interactions between the region of Bcl-2 where BH1 and BH2 resides, and an NH-terminal domain in Bcl-2 that contains another domain BH4 which is conserved among antiapoptotic members of the Bcl-2 family. Similarly, heterodimerization with Bcl-2 occurred without the NH-terminal domain of either Bax or Bcl-2, suggesting a tail-to-tail interaction. The essential region in Bax required for both homodimerization with Bax and heterodimerization with Bcl-2 was mapped to residues 59-101. This region in Bax contains a stretch of 15 amino acids that is highly homologous in several members of the Bcl-2 protein family, suggesting the existence of a novel functional domain which we have termed BH3. Deletion of this 15-amino acid region abolished the ability of Bax to dimerize with itself and to heterodimerize with Bcl-2. The findings suggest that the structural features of Bax and Bcl-2 that allow them to participate in homo- and heterodimerization phenomena are markedly different, despite their amino-acid sequence similarity.
Article
Full-text available
Bax is a Bcl-2-family protein with pro-apoptotic activity that can form channels in lipid membranes. The protein has been shown to trigger cytochrome c release from mitochondria both in vitro and in vivo. Recombinant human Bax isolated in the presence of detergent was found to be present as an oligomer with an apparent molecular mass of approx. 160000 Da on gel filtration. When Bax was isolated in the absence of detergent the purified protein was monomeric with an apparent molecular mass of 22000 Da. Bax oligomers formed channels in liposomes and triggered cytochrome c release from isolated mitochondria, whereas monomeric Bax was inactive in both respects. Incubation of the monomeric Bax with 2% octyl glucoside induced formation of oligomers that displayed channel-forming activity in liposomes and triggered cytochrome c release from mitochondria. Triton X-100, Nonidet P-40 and n-dedecyl maltoside also activated monomeric Bax, whereas CHAPS had no activating effect. In cytosolic extracts from mouse liver, Bax migrated at a molecular mass of 24000 Da on gel filtration, whereas after incubation of the cytosol with 2% octyl glucoside Bax migrated at approximately 140000 Da. These results show that oligomeric Bax possesses channel-forming activity whereas monomeric Bax has no such activity.
Article
Most members of the Bcl-2 protein family of apoptosis regulating proteins contain two evolutionarily conserved domains, termed BH1 and BH2. Both BH1 and BH2 in the Bcl-2 protein are required for its function as an inhibitor of cell death and for heterodimerization with the proapoptotic protein Bax. In this report, we mapped the region in Bax required for heterodimerization with Bcl-2 and homodimerization with Bax, using yeast two-hybrid and in vitro protein-protein interaction assays. Neither the BH1 nor the BH2 domain of Bax was required for binding to the wild-type Bcl-2 and Bax proteins. Moreover, Bax (Delta BH1) and Bax (Delta BH2) mutant proteins bound efficiently to themselves and each other, further confirming the lack of requirement for BH1 and BH2 for Bax/Bax homodimerization. Bax/Bax homodimerization was not dependent on the inclusion of the NH2-terminal 58 amino acids of the Bax protein in each dimerization partner, unlike Bcl-2/Bcl-2 homodimers which involve head-to-tail interactions between the region of Bcl-2 where BH1 and BH2 resides, and an NH2-terminal domain in Bcl-2 that contains another domain BH4 which is conserved among antiapoptotic members of the Bcl-2 family. Similarly, heterodimerization with Bcl-2 occurred without the NH2-terminal domain of either Bax or Bcl-2, suggesting a tail-to-tail interaction. The essential region in Bax required for both homodimerization with Bax and heterodimerization with Bcl-2 was mapped to residues 59-101. This region in Bax contains a stretch of 15 amino acids that is highly homologous in several members of the Bcl-2 protein family, suggesting the existence of a novel functional domain which we have termed BH3. Deletion of this 15-amino acid region abolished the ability of Bax to dimerize with itself and to heterodimerize with Bcl-2. The findings suggest that the structural features of Bax and Bcl-2 that allow them to participate in homo- and heterodimerization phenomena are markedly different, despite their amino-acid sequence similarity.
Article
Extracellular survival factors alter a cell's susceptibility to apoptosis, often through posttranslational mechanisms. However, no consistent relationship has been established between such survival signals and the BCL-2 family, where the balance of death agonists versus antagonists determines susceptibility. One distant member, BAD, heterodimerizes with BCL-X(L) or BCL-2, neutralizing their protective effect and promoting cell death. In the presence of survival factor IL-3, cells phosphorylated BAD on two serine residues embedded in 14-3-3 consensus binding sites. Only the nonphosphorylated BAD heterodimerized with BCL-X(L) at membrane sites to promote cell death. Phosphorylated BAD was sequestered in the cytosol bound to 14-3-3. Substitution of serine phosphorylation sites further enhanced BAD's death-promoting activity. The rapid phosphorylation of BAD following IL-3 connects a proximal survival signal with the BCL-2 family, modulating this checkpoint for apoptosis.
Article
Mast cells reside in tissues, where upon activation through the high-affinity-IgE-receptor (FcRI) they degranulate and orchestrate the allergic reaction. Mast cells survive this activation and can thus be reactivated. In this study we demonstrate that this process depends on the pro- survival gene A1. Activation of mast cells through FcRI resulted in degranulation, strong in- duction of A1 mRNA and protein, and cell survival. In contrast, A1-deficient mast cells released granule mediators similar to the wild-type control, but the cells did not survive an al- lergic activation. Furthermore, A1 � / � mice that had been sensitized and provocated with aller- gen exhibited a lower number of mast cell compared with littermate controls. The induction of A1 was dependent on calcium, as EDTA prevented A1 expression. The calcium ionophore, iono- mycin, induced A1 expression and mast cell survival, whereas compound 48/80, a well-known mast cell secretagogue, did not. This study uncovers the importance of A1 for mast cell survival in allergic reactions, and it proposes A1 as a potential target for the treatment of allergic diseases.
Article
Using a Bax-dependent membrane-permeabilization assay, we show that peptides corresponding to the BH3 domains of Bcl-2 family “BH3-only” proteins have dual functions. Several BH3 peptides relieved the inhibition of Bax caused by the antiapoptotic Bcl-xL and/or Mcl-1 proteins, some displaying a specificity for either Bcl-xL or Mcl-1. Besides having this derepression function, the Bid and Bim peptides activated Bax directly and were the only BH3 peptides tested that could potently induce cytochrome c release from mitochondria in cultured cells. Furthermore, Bax activator molecules (cleaved Bid protein and the Bim BH3 peptide) synergistically induced cytochrome c release when introduced into cells along with derepressor BH3 peptides. These observations support a unified model of BH3 domain function, encompassing both positive and negative regulation of other Bcl-2 family members. In this model, the simple inhibition of antiapoptotic functions is insufficient to induce apoptosis unless a direct activator of Bax or Bak is present.
Article
The overexpression of Bax, a member of the Bcl-2 family, promotes cell death and the dimerization (or oligomerization) of Bax has been shown to be important for its function. Using size-exclusion chromatography andin vitrocross-linking experiments, we demonstrated that Bax exists mainly as a large oligomer of approximately 30 monomeric units. Furthermore, several binding assays demonstrated that Bcl-XL, an anti-apoptotic member of the Bcl-2 family, can bind to the oligomeric form of Bax without requiring Bax to dissociate to monomers.