Under stress conditions, apoptogenic factors normally sequestered in the mitochondrial intermembrane space are released into the cytosol, caspases are activated and cells die by apoptosis. Although the precise mechanism that leads to the permeabilization of mitochondria is still unclear, the activation of multidomain pro-apoptotic proteins of the Bcl-2 family, such as Bax and Bak, is evidently crucial. Regulation of Bax and Bak by other members of the family has been known for a long time, but recent evidence suggests that additional unrelated proteins participate in the process, both as inhibitors and activators. The important rearrangements mitochondrial lipids undergo during apoptosis play a role in the permeabilization process and this role is probably more central than first envisioned.

Programmed cell death is essential for the development of multicellular organisms, for tissue homeostasis, and to eliminate damaged and infected cells. Apoptosis is one form of programmed cell death. It is defined by specific biochemical and morphological changes, including DNA fragmentation, compaction of the chromatin into dense structures, blebbing of the plasma membrane, exposure of phosphatidylserine in the outer leaflet of the plasma membrane, and engulfment of the apoptotic bodies by phagocytes. These changes follow the activation of the cysteinyl aspartate-specific proteinases known as caspases (Fischer et al., 2003; Nicholson, 1999).

Two main pathways leading to the activation of caspases have been identified (Zimmermann et al., 2001). The first, known as the extrinsic pathway, is activated when ligands bind to death receptors on the plasma membrane. The second, the intrinsic pathway, is activated by various stress signals, such as DNA damage, growth factor withdrawal, or anoikis (cell detachment), and involves the permeabilization of mitochondria, which then release apoptogenic factors (e.g. cytochrome c, Smac/DIABLO and Omi/HtrA2) normally sequestered in the mitochondrial intermembrane space. Crosstalk can occur between these two pathways, as found in some cell types in which death-receptor-induced apoptosis involves mitochondria (Scaffidi et al., 1998). However, these are not the only pathways; recent work has identified alternative mechanisms for the activation of caspase-2 (Lassus et al., 2002), caspase-4 (Hitomi et al., 2004) and caspase-12 (Morishima et al., 2002; Nakagawa et al., 2000).

Permeabilization of mitochondria is regulated by proteins of the Bcl-2 family, which are defined by the presence of at least one of four Bcl-2-homology domains (BH1 to BH4) (Borner, 2003; Cory et al., 2003; Willis et al., 2003). Bcl-2 family proteins have been subdivided into three groups on the basis of their pro- or anti-apoptotic action and the BH domains they possess. Anti-apoptotic Bcl-2-like proteins (e.g. Bcl-2, Bcl-xL, Bcl-w, Mcl-1 and A1/Bfl-1) have the four BH domains; pro-apoptotic multidomain proteins (e.g. Bax, Bak and Bok/Mtd) lack BH4; and, as their name indicates, pro-apoptotic BH3-only proteins (e.g. Bid, Bim/Bod, Bad, Bmf, Bik/Nbk, Blk, Noxa, Puma/Bbc3 and Hrk/DP5) possess only a BH3 domain. Some proteins of the Bcl-2 family have in addition a hydrophobic C-terminal tail that allows their association with membranes (Schinzel et al., 2004a). The level of expression of each member of the family is tightly regulated and post-translational modifications modulate their activities (Borner, 2003; Willis et al., 2003). Although their mechanisms of action are incompletely understood, heterodimerization between pro- and anti-apoptotic Bcl-2-like proteins is known to be important. Structural studies have revealed that a hydrophobic groove on the surface of anti-apoptotic Bcl-2 family members is the binding site for the BH3 domain of their pro-apoptotic counterparts (Petros et al., 2004).

Pro-apoptotic multidomain proteins seem to be the active members of the Bcl-2 family that trigger the release of mitochondrial apoptogenic factors into the cytosol. The presence of Bak or Bax is known to be crucial for numerous stress signals to induce cell death (Wei et al., 2001). By contrast, Bok/Mtd has been less examined, perhaps because of its restricted tissue distribution and its small number of known binding partners among Bcl-2 family members (Hsu et al., 1997a; Inohara et al., 1998). From extensive studies on the sequence of conformational changes that Bak and Bax undergo when they are activated, the following model has emerged (Fig. 1). In resting cells, Bak is an integral protein of the mitochondrial outer membrane, whereas Bax is cytosolic or loosely associated with the mitochondria in a form that can be detached from membranes by alkali treatment (Desagher et al., 1999; Goping et al., 1998; Hsu et al., 1997b). In response to certain apoptotic signals, the conformations of Bax and Bak change. Their N-terminal domains, which are not accessible to antibodies in the inactive proteins, become exposed, and Bax translocates to the mitochondria and integrates into the outer membrane as an alkali-resistant form (Desagher et al., 1999; Goping et al., 1998; Hsu et al., 1997b; Hsu and Youle, 1997; Makin et al., 2001). The subsequent oligomerization of Bak and/or Bax is thought to result in the permeabilization of the mitochondrial outer membrane (Antonsson et al., 2000).

Fig. 1.

Model for the mechanism of activation of Bak and Bax. In resting cells, Bak is a tail-anchored protein of the mitochondrial outer membrane (MOM) and associates with VDAC2, whereas Bax is loosely attached to mitochondria or sequestered in the cytosol, probably through interactions with retention factors (14-3-3 isoforms σ, θ, ϵ, ζ; humanin; Ku70; αA- and αB-crystallin; Hsp70-dj1 and Hsp70-dj2). Pro168 plays a crucial role both in preventing the inappropriate exposure of the N-terminal domain of Bax and in unleashing the C-terminal tail of Bax from its hydrophobic pocket as it occurs when Bax is activated (Schinzel et al., 2004b). After apoptosis induction, the conformations of Bak and Bax change, leading to exposure of their N-terminal domains and to oligomerization. Apoptogenic factors are then released from the mitochondrial intermembrane space. Several proteins have been suggested to participate in the activation of Bak and Bax, although the mechanisms through which they contribute remain unclear. In addition, some proteins promote apoptosis by inhibiting Bcl-2-like proteins.

Fig. 1.

Model for the mechanism of activation of Bak and Bax. In resting cells, Bak is a tail-anchored protein of the mitochondrial outer membrane (MOM) and associates with VDAC2, whereas Bax is loosely attached to mitochondria or sequestered in the cytosol, probably through interactions with retention factors (14-3-3 isoforms σ, θ, ϵ, ζ; humanin; Ku70; αA- and αB-crystallin; Hsp70-dj1 and Hsp70-dj2). Pro168 plays a crucial role both in preventing the inappropriate exposure of the N-terminal domain of Bax and in unleashing the C-terminal tail of Bax from its hydrophobic pocket as it occurs when Bax is activated (Schinzel et al., 2004b). After apoptosis induction, the conformations of Bak and Bax change, leading to exposure of their N-terminal domains and to oligomerization. Apoptogenic factors are then released from the mitochondrial intermembrane space. Several proteins have been suggested to participate in the activation of Bak and Bax, although the mechanisms through which they contribute remain unclear. In addition, some proteins promote apoptosis by inhibiting Bcl-2-like proteins.

Insertion of Bax into the mitochondria requires the release of its C-terminal hydrophobic tail from the hydrophobic pocket formed by its BH1, BH2 and BH3 domains (Nechushtan et al., 1999; Schinzel et al., 2004b). Deletion of proline 168 (Pro168) abolishes the capacity of Bax to translocate after a cytotoxic insult (Schinzel et al., 2004b). Isomerization or post-translational modification of Pro168 might therefore be required to disrupt the intramolecular interactions that maintain Bax as a soluble protein. Alternatively, Pro168 could be essential for the association of Bax with specific regulators (Schinzel et al., 2004b).

BH3-only proteins are crucial intermediates that link specific apoptotic stimuli to the permeabilization of mitochondria (Puthalakath and Strasser, 2002). Even though killing by BH3-only proteins requires Bax or Bak, most of them seem to promote apoptosis by binding to and inhibiting pro-survival Bcl-2 proteins (Fig. 2A) (Cheng et al., 2001). Only Bid (Desagher et al., 1999; Eskes et al., 2000; Wei et al., 2000), some Bim isoforms (Marani et al., 2002) and the distantly related protein MAP-1 (Tan et al., 2001) have been suggested to bind to Bax and Bak and induce their activation. Incubation of isolated mitochondria with recombinant Bax and tBid, which is the active form of Bid generated by proteolytic cleavage, leads to the oligomerization of Bax and to cytochrome c release (Desagher et al., 1999; Eskes et al., 2000). However, Bax cannot be activated if mitochondria are pre-incubated with proteinase K, which indicates that additional factors are required for tBid to activate Bax (Roucou et al., 2002). Analysis of the 3D structure of soluble Bax has revealed that binding of the BH3 domain of another Bcl-2 family member would not be enough to disrupt the interactions between the C-terminal tail of Bax and its hydrophobic pocket (Suzuki et al., 2000). However, some reports suggest that, in the presence of pure synthetic liposomes, tBid could be sufficient (Kuwana et al., 2002; Terrones et al., 2004).

Fig. 2.

(A) Following an apoptotic insult, BH3-only proteins are activated, at the transcriptional and/or the post-translational level, and neutralize their pro-survival counterparts. Activation of multidomain pro-apoptotic members of the Bcl-2 family proceeds through a largely unknown mechanism, although some BH3-only factors have been proposed to participate in the process. (B) Under resting conditions, pro-survival Bcl-2 family members sequester low levels of BH3-only proteins and prevent the full activation of multidomain Bax-like factors that would have undergone conformational rearrangements resulting in the exposure of their N-terminal domains.

Fig. 2.

(A) Following an apoptotic insult, BH3-only proteins are activated, at the transcriptional and/or the post-translational level, and neutralize their pro-survival counterparts. Activation of multidomain pro-apoptotic members of the Bcl-2 family proceeds through a largely unknown mechanism, although some BH3-only factors have been proposed to participate in the process. (B) Under resting conditions, pro-survival Bcl-2 family members sequester low levels of BH3-only proteins and prevent the full activation of multidomain Bax-like factors that would have undergone conformational rearrangements resulting in the exposure of their N-terminal domains.

Every mammalian cell type seems to require the expression of at least one anti-apoptotic Bcl-2-like protein to survive (Ranger et al., 2001). These proteins can protect cells from apoptosis at multiple levels (Cory et al., 2003; Gross et al., 1999). They can sequester BH3-only factors (Fig. 2B) (Cheng et al., 2001), and even though they do not appear to bind to inactive Bax and Bak, they can neutralize them if the latter have undergone conformational changes resulting in the exposure of their N-terminal domains, preventing their oligomerization and the release of apoptogenic factors from mitochondria (Desagher et al., 1999; Hsu and Youle, 1997; Perez and White, 2000; Ruffolo and Shore, 2003). Indeed, heterodimerization only occurs when non-ionic detergents, which artificially promote rearrangements that Bax and Bak normally undergo during the intrinsic pathway of apoptosis, are included in the extracts (Hsu and Youle, 1997). The pro-survival actions of Bcl-2-like proteins extend beyond the direct inhibition of pro-apoptotic members of the family, as they seem to control endoplasmic reticulum (ER) and mitochondrial homeostasis, probably by regulating Ca2+ fluxes (Distelhorst and Shore, 2004) and by mechanisms that defend against oxidative stress (Jang and Surh, 2003).

In addition to members of the Bcl-2 family, numerous proteins have been identified as Bax- and/or Bak-interacting proteins that prevent or promote the conformational rearrangements that Bax and Bak undergo during apoptosis. Below, we focus on these new actors and discuss the roles that lipids have recently been suggested to play in the permeabilization of mitochondria.

Proteins that inhibit the activation of Bax and Bak

Several proteins have been reported to interact with Bax under resting conditions. Gel filtration analysis suggests that the soluble form of Bax is monomeric (Antonsson et al., 2000), but it might associate in vivo with cytosolic retention factors (Fig. 1). 14-3-3 isoforms (σ, θ, ϵ, ζ), the small peptide humanin, Ku70 and the apoptosis repressor with caspase-recruitment domain (ARC) were identified as Bax-interacting proteins whose overexpression inhibits Bax-dependent apoptosis, preventing its translocation to mitochondria (Guo et al., 2003; Gustafsson et al., 2004; Nam et al., 2004; Nomura et al., 2003; Samuel et al., 2001; Sawada et al., 2003). Similarly, Bax is retained in the cytosol when cells stably transfected with the Bax-interacting heat shock proteins (Hsps) αA- and αB-crystallin or the chaperone pairs Hsp70-dj1 and Hsp70-dj2 are exposed to stimuli that normally promote the intrinsic pathway of apoptosis (Gotoh et al., 2004; Mao et al., 2004). Knocking down humanin or Ku70 by RNA interference (RNAi) and knocking out 14-3-3σ only sensitizes cells to Bax-dependent apoptosis (Guo et al., 2003; Samuel et al., 2001; Sawada et al., 2003), whereas downregulation of ARC results in spontaneous exposure of the N-terminal domain of Bax and perhaps to increased cell death (Nam et al., 2004).

A more thorough examination of the properties of these proteins will be required to determine whether the binding of Bax is important for their anti-apoptotic actions. Proteins of the 14-3-3 family are adaptors that modulate the localization and activities of a wide array of targets and are known to have other anti-apoptotic activities - for example, through the sequestration of the BH3-only protein Bad - and to inhibit cell-cycle progression following DNA damage (Dougherty and Morrison, 2004). Therefore, modulating 14-3-3 levels will certainly affect several mechanisms that contribute to cell death and the effects observed cannot be unequivocally associated with Bax binding. Humanin was initially identified as a secreted peptide that has a neuroprotective effect, but under certain conditions it might bind to Bax in the cytosol and prevent its translocation to mitochondria (Hashimoto et al., 2001). The Ku70-Ku80 complex is involved in DNA double-strand repair. Although overexpression of Ku80 does not suppress Bax-mediated apoptosis (Sawada et al., 2003), the phenotypes of Ku70-/- mice and of Ku80-/- mice are similar (Gu et al., 1997). This suggests that, even though a cytosolic fraction of Ku70 associates with Bax, it might not be relevant under physiological conditions.

αA- and αB-crystallin protect lens epithelial cells from cell death induced by a variety of apoptotic insults (Andley et al., 2000). However, in addition to Bax sequestration, two other protection mechanisms have been proposed: inhibition of pro-caspase-3 processing and inhibition of the upregulation of some pro-apoptotic genes (Kamradt et al., 2002; Mao et al., 2004). Because αA- and αB-crystallin are primarily expressed in the lens and the phenotypes of knockouts do not suggest abnormal activation of apoptotic pathways (Brady et al., 1997; Brady et al., 2001), their Bax-binding activity does not appear to be a general mechanism of regulation of Bax. Hsp70-dj1 and Hsp70-dj2 complexes prevent Bax translocating to mitochondria after ER-stress-induced apoptosis. Hsp70 has been proposed to inhibit cell death at multiple levels and Hsp70 knockouts are characterized by genomic instability and increased radiosensitivity (Hunt et al., 2004). Therefore, the importance of the sequestration of Bax by these complexes will need to be examined further.

ARC is the only Bax-interacting protein whose downregulation results in exposure of the N-terminus of Bax and increased cell death (Nam et al., 2004). However, the insertion of Bax into the mitochondria and its oligomerization status have not been assessed, and cell death was only monitored by trypan blue exclusion in these experiments. Because ARC interacts with other proteins involved in apoptosis regulation, such as death receptors and adaptors, procaspase-2 and procaspase-8, the relevance of its association with Bax for its anti-apoptotic actions still needs to be confirmed (Koseki et al., 1998; Nam et al., 2004).

Another potential mechanism of inhibition of Bax mitochondrial translocation involves hexokinase II. Overexpression of hexokinase II, a glycolytic enzyme associated with the mitochondrial outer membrane, also prevents the translocation of Bax to mitochondria (Pastorino et al., 2002). The overexpressed hexokinase might block access of Bax to contact sites that are already highly enriched in proteins (Majewski et al., 2004).

Voltage-dependent anion channel 2 (VDAC2), which is a relative of the VDAC protein perhaps implicated in mitochondrial permeabilization (see below), was recently identified as a Bak-interacting protein that stabilizes Bak in an inactive conformation (Cheng et al., 2003). Downregulation of VDAC2 sensitizes cells to multiple apoptotic stimuli, and VDAC2-/- mouse embryonic fibroblasts are more susceptible to apoptosis than are wild-type cells (Cheng et al., 2003).

Proteins that stimulate the activation of Bax and Bak

In addition to the proteins that might stabilize Bax and Bak in inactive conformations, others have been proposed to promote their activation (Fig. 1). Bax associates with endophilin B1a (also known as Bif-1) in FL5.12 cells, and the degree of interaction increases after withdrawal of interleukin 3 (Cuddeback et al., 2001). Endophilin B1a translocates to mitochondria during apoptosis (Karbowski et al., 2004) and its overexpression accelerates apoptosis under cytotoxic conditions (Cuddeback et al., 2001). Endophilin B1a is similar to endophilin A1, a protein involved in the regulation of synaptic vesicle endocytosis that binds to components of the endocytosis machinery and can modify phospholipids in the bilayer (Huttner and Schmidt, 2002). Similarly, downregulation of endophilin B1a by RNAi alters mitochondrial morphology (Karbowski et al., 2004). Endophilin B1a might contribute to the permeabilization of mitochondria by modifying the properties of the lipid bilayer or by bridging Bax to components of the mitochondrial fission machinery. Indeed, during apoptosis, mitochondria fragment and cluster in the perinuclear region (Desagher and Martinou, 2000; Frank et al., 2001), and components of the mitochondrial fusion and fission apparatus are thought to modulate their permeabilization (Karbowski and Youle, 2003).

The p53 tumor suppressor accumulates in cells exposed to various stress conditions and can either promote growth arrest or apoptosis (Slee et al., 2004). In addition to its effects on gene expression, which sometimes seem to be essential for apoptosis (Haupt et al., 2003), p53 can also promote cell death through transcription-independent mechanisms (Fig. 3) (Chipuk and Green, 2003; Chipuk et al., 2004; Slee et al., 2004). Following certain cytotoxic signals, a small fraction of p53 translocates to mitochondria (Erster et al., 2004; Marchenko et al., 2000). Interaction of p53 with Bcl-2 and Bcl-xL (Chipuk et al., 2004; Mihara et al., 2003), and with Bak (Leu et al., 2004), has been reported. However, although cytochrome c release has been observed by some groups when they incubate isolated mitochondria with purified p53 (Leu et al., 2004; Mihara et al., 2003), another group failed to see any release unless Bax is also included (Chipuk et al., 2004). Therefore, it is unclear whether p53 promotes the permeabilization of mitochondria by inhibiting pro-survival Bcl-2-like proteins and/or by activating Bak or Bax. p53 might also lead to caspase activation through mitochondria-independent pathways (Ding et al., 1998), perhaps by promoting the activation of caspase-2 (Lin et al., 2000; Tinel and Tschopp, 2004).

Fig. 3.

Transcription-dependent and -independent mechanisms of p53-mediated apoptosis. In a cell-context-specific manner, p53 can activate the expression of pro-apoptotic genes and repress the transcription of pro-survival proteins (Ho and Benchimol, 2003). p53 can also translocate to the cytosol and associate with mitochondria. Direct effects of p53 on proteins of the Bcl-2 family have been proposed, as well as indirect effects through the upregulation of BH3-only proteins, p53AIP1 and ASC. In addition, p53 can promote the production of reactive oxygen species (ROS; Li et al., 1999) and activate caspases independently of mitochondria, perhaps through upregulation of PIDD (Lin et al., 2000), a protein involved in the activation of caspase-2 (Tinel and Tschopp, 2004).

Fig. 3.

Transcription-dependent and -independent mechanisms of p53-mediated apoptosis. In a cell-context-specific manner, p53 can activate the expression of pro-apoptotic genes and repress the transcription of pro-survival proteins (Ho and Benchimol, 2003). p53 can also translocate to the cytosol and associate with mitochondria. Direct effects of p53 on proteins of the Bcl-2 family have been proposed, as well as indirect effects through the upregulation of BH3-only proteins, p53AIP1 and ASC. In addition, p53 can promote the production of reactive oxygen species (ROS; Li et al., 1999) and activate caspases independently of mitochondria, perhaps through upregulation of PIDD (Lin et al., 2000), a protein involved in the activation of caspase-2 (Tinel and Tschopp, 2004).

Upon DNA damage, p53 also upregulates the expression of several pro-apoptotic proteins that localize to mitochondria (Fig. 3). Among these, p53-regulated apoptosis-inducing protein 1 (p53AIP1) can interact with Bcl-2 (Matsuda et al., 2002), and the apoptosis-associated speck-like protein (ASC) can interact with Bax (Ohtsuka et al., 2004). Overexpression of either protein leads to cell death, and their downregulation inhibits apoptosis following DNA damage (Matsuda et al., 2002; Oda et al., 2000). ASC can also lead to caspase activation and apoptosis independently of Bax and Bak (Ohtsuka et al., 2004) and was recently shown to be important for maturation of the inflammatory caspase-1 (Mariathasan et al., 2004). ASC-deficient cells will need to be further tested in order to analyze their sensitivity to DNA-damage-induced cell death.

Under certain apoptotic conditions, pro-survival proteins of the Bcl-2 family seem not only to be neutralized but also even to become pro-apoptotic. For example, removal of the BH4 domains of Bcl-2 and Bcl-xL by caspase cleavage makes them apoptogenic (Cheng et al., 1997; Clem et al., 1998). Following certain apoptotic signals, the nuclear orphan receptor Nur77/TR3 translocates from the nucleus to mitochondria and promotes cytochrome c release (Li et al., 2000). It binds to Bcl-2, inducing the exposure of its BH3 domain and reducing the accessibility of its hydrophobic pocket (Lin et al., 2004). By interacting with Nur77/TR3, Bcl-2 becomes pro-apoptotic, but still requires Bak to lead to cell death (Lin et al., 2004). Even though Nur77 knockouts have a minimal phenotype (Lee et al., 1995), Nur77/TR3 could participate in T-cell apoptosis (Calnan et al., 1995). Recently, Kim et al. reported that Bcl-2 changes its membrane topology during apoptosis (Kim et al., 2004a). Although this could be required for the anti-apoptotic activity of Bcl-2, it might represent a mechanism through which Bcl-2 is converted to a pro-apoptotic factor (Puthalakath and Strasser, 2002).

Several other proteins can promote the permeabilization of mitochondria, but it is still unclear whether they also regulate members of the Bcl-2 family. The linker histone H1 is released into the cytosol when cells are treated with DNA-damaging agents that cause double-strand breaks (Konishi et al., 2003). The subtype H1.2 induces the activation of Bak and release of cytochrome c, but no direct interaction with Bcl-2 proteins has been detected (Konishi et al., 2003). After apoptosis induction, the actin-depolymerizing factor cofilin rapidly translocates from the cytosol to mitochondria (Chua et al., 2003). Downregulation of cofilin prevents apoptosis, whereas the overexpression of a cofilin mutant that constitutively localizes to mitochondria leads to cell death (Chua et al., 2003). Finally, the expression level of the chloride intracellular channel mtCLIC/CLIC4 increases under several apoptotic conditions (Fernandez-Salas et al., 1999; Fernandez-Salas et al., 2002). Cell death that follows the overexpression of mtCLIC/CLIC4 is independent of Bax but is inhibited by Bcl-2 (Suh et al., 2004), and downregulation of mtCLIC/CLIC4 inhibits p53-induced cell death (Fernandez-Salas et al., 2002). mtCLIC/CLIC4 could promote classical intrinsic apoptotic pathways by modifying intra-organelle Cl- levels, but it remains possible that at the mitochondria it directly promotes cytochrome c release.

Relevance of these interactions

Most of the putative Bax and Bak regulators have been identified by co-immunoprecipitation, and their relevance to cell death has been proposed from studies showing modulation of the sensitivity to apoptotic insults following their overexpression or their downregulation by RNAi. Such results are important, but should be interpreted with caution. For example, the multiplicity of Bcl-2-binding partners could be explained by the stickiness of the Bcl-2-like proteins. It is also difficult to understand how downregulation of each putative Bax-binding protein always results in approximately 30% inhibition of, or sensitization to, apoptosis. Indeed, it is conceivable that downregulation of a bona fide Bax-binding protein might have no impact on cell death, but such results have so far never been reported. Even if they do not allow the resolution of all the issues regarding the roles of these Bax regulators, a thorough analysis of knockouts and their sensitivity to apoptotic insults, together with a detailed study of the conformations of Bax and Bak, will be required to determine precisely which of these proteins, if any, are relevant apoptosis regulators.

The precise mechanism that leads to the permeabilization of mitochondria is still unclear, despite several models having been extensively debated (Crompton et al., 2002; Desagher and Martinou, 2000; Martinou and Green, 2001; Waterhouse et al., 2002). The first invokes opening of permeability transition (PT) pores at contact sites between the inner and outer mitochondrial membranes, subsequent swelling of the matrix and rupture of the outer membrane, which has a smaller surface area than the inner membrane. VDAC, the adenine-nucleotide translocator (ANT) and the matrix chaperone cyclophilin D were initially proposed to be central components of these pores (Crompton et al., 1998; Halestrap et al., 2002), and Bcl-2 family proteins were proposed to regulate their opening (Zamzami and Kroemer, 2001). Inhibition of cytochrome c release by the cyclophilin-D-binding drug cyclosporin A (CsA) supports this model. However, CsA is not as specific as initially thought, and binds not only to all the members of the cyclophilin family but also to other proteins (Fruman et al., 1992; Moss et al., 1992). Moreover, it partitions in membranes and could affect their permeability indirectly by modifying the properties of the lipid bilayer (Epand et al., 2002a; McGuinness et al., 1990). Furthermore, this model has recently been challenged by the observation that ANT-deficient mitochondria can still undergo PTs and ANT-deficient cells can die by apoptosis (Kokoszka et al., 2004).

An alternative model is that modulation of the properties of VDAC by Bcl-2 proteins is sufficient to control the permeability of mitochondria (Tsujimoto and Shimizu, 2000). Finally, given their structural similarity to some bacterial toxins, Bcl-2 family proteins might form proteinaceous channels themselves (Epand et al., 2002c; Saito et al., 2000). These models might reflect the events that occur during apoptosis but are incomplete, because it is increasingly clear that lipids have an important role in the permeabilization of mitochondria.

Lipids and apoptosis

During apoptosis, lipid membranes undergo important rearrangements (Cristea and Degli Esposti, 2004; Wright et al., 2004). Exposure of phosphatidylserine on the external face of the plasma membrane and plasma membrane blebbing were the first of these to be discovered. More recently, several groups reported that the levels of cardiolipin diminish when cell death is induced by a variety of conditions (Matsko et al., 2001; Nomura et al., 2000; Ostrander et al., 2001). Cardiolipin is a negatively charged phospholipid that mainly resides in the mitochondrial inner membrane, where it is required for the activities of several proteins of the electron transport chain, including cytochrome c, and of several mitochondrial carriers, including ANT (McMillin and Dowhan, 2002). Several groups have reported that peroxidation of cardiolipin occurs during apoptosis (Garcia Fernandez et al., 2002; Nomura et al., 2000) and is required for detachment of cytochrome c from the mitochondrial inner membrane and its complete release (Nomura et al., 2000; Ott et al., 2002; Petrosillo et al., 2001; Shidoji et al., 1999).

Cardiolipin also appears to be present in the mitochondrial outer membrane, enriched at contact sites with the inner membrane, where it could interact with Bcl-2 family proteins (Ardail et al., 1990; Hovius et al., 1993; Qi et al., 2003). Indeed, activated Bax seems to permeabilize synthetic liposomes only if cardiolipin is present (Kuwana et al., 2002; Terrones et al., 2004). Interestingly, the number and the surface area of contact sites increase in rod photoreceptors of mice exposed to toxic insults (He et al., 2003), and oligomeric Bax lacking its C-terminal tail (BaxΔC) does not permeabilize outer membrane vesicles if the contact sites have been removed (Wieckowski et al., 2001). These results highlight the significance that the specific lipid and/or protein composition of contact sites might have for the action of Bax and Bak.

Upon activation, tBid is recruited to mitochondria, in particular to contact sites, where it binds to cardiolipin or its derivative monolysocardiolipin, produced during Fas-induced cell death (Lutter et al., 2000; Sorice et al., 2004). When mouse liver mitochondria are incubated with tBid and when cells undergo apoptosis, mitochondrial cristae remodel (Scorrano et al., 2002). They become more interconnected and cristae junctions widen. CsA inhibits remodeling (Scorrano et al., 2002), either because CsA targets a protein involved in the process or because of its direct effects on membrane curvature and stability (Epand et al., 2002a; Epand et al., 1987). Pre-incubation of mitochondria with the cardiolipin-specific dye 10-N-nonyl acridine orange (NAO) prevents the binding of tBid to contact sites, and cristae remodeling and cytochrome c release are abolished, even though Bak still oligomerizes (Kim et al., 2004b). Such cristae opening might increase the pool of cytochrome c available for release. However, association of tBid with contact sites might also be directly necessary for the complete permeabilization of mitochondria by activated Bax or Bak.

Oligomeric Bax, which stimulates transbilayer lipid diffusion (Epand et al., 2003; Terrones et al., 2004), and tBid, which possesses a lipid transfer activity (Degli Esposti, 2002; Esposti et al., 2001), might also contribute to the redistribution of cardiolipin between the mitochondrial membranes, to intracellular organelles and to the plasma membrane, as seems to occur during apoptosis (Qi et al., 2003; Sorice et al., 2004). A rat liver fatty-acid-binding protein released from isolated mitochondria after incubation with tBid could also contribute (Van Loo et al., 2002).

Other studies have focused on the influence of the composition of liposomes on their permeabilization by activated Bax. Although a particular lipidic composition might be required only for oligomeric Bax to adopt an active conformation, lipids seem to play an active part in the permeabilization process. Basañez and coworkers reported that this is potentiated by lipids that have a positive intrinsic curvature, such as lysophospholipids [oleoyl-phosphatidylcholine (LPC) and oleoyl-phosphatidyl-ethanolamine (LPE)], and is inhibited by lipids that have a negative intrinsic curvature, such as dioleoylglycerol (DOG) or dioleoylphosphatidylethanolamine (DOPE) (Basañez et al., 2002; Terrones et al., 2004) (Fig. 4). Moreover, oligomeric Bax destabilizes planar lipid bilayers, such that they ultimately rupture, and this effect is similarly modulated by the presence of the above-mentioned lipids (Basañez et al., 1999; Basañez et al., 2002). In addition, both LPC and LPG increase the amount of cytochrome c that is released upon incubation of mouse liver mitochondria with tBid (Degli Esposti, 2002; Esposti et al., 2003). This behavior is reminiscent of the bacterial pore-forming peptide magainin (Matsuzaki et al., 1998) and suggests that the mechanism of Bax-induced permeabilization of synthetic liposomes involves lipid pores (Fig. 4). The Bax-interacting protein endophilin B1a binds to fatty acids and possesses a lysophosphatidic acid acyltransferase activity similar to that of endophilin A1, which participates in synaptic vesicle formation by increasing membrane positive curvature (Modregger et al., 2003). Endophilin B1a could similarly help Bax create lipid pores in the mitochondrial outer membrane.

Fig. 4.

When lipids are assimilated with solid structures, they can be divided into three groups (Israelachvili and Mitchell, 1975): flat lipids, which are roughly cylindrical; lipids with a positive curvature, which have a wider hydrophilic headgroup than the cross-sectional surface occupied by their acyl chains; and lipids with negative curvature, which have a smaller headgroup than the area of a cross-section of their hydrophobic part. Accumulation of lipids with positive curvature can create lipid pores in a lamellar bilayer, whereas lipids with negative curvature adopt a non-lamellar structure called hexagonal II phase. Both of these structures have been suggested to contribute to the permeabilization of the mitochondrial outer membrane that occurs during apoptosis. PC, phosphatidylcholine; PG, phosphatidylglycerol; PS, phosphatidylserine; PI, phosphatidylinositol; PE, phosphatidylethanolamine; DAG, diacylglycerol.

Fig. 4.

When lipids are assimilated with solid structures, they can be divided into three groups (Israelachvili and Mitchell, 1975): flat lipids, which are roughly cylindrical; lipids with a positive curvature, which have a wider hydrophilic headgroup than the cross-sectional surface occupied by their acyl chains; and lipids with negative curvature, which have a smaller headgroup than the area of a cross-section of their hydrophobic part. Accumulation of lipids with positive curvature can create lipid pores in a lamellar bilayer, whereas lipids with negative curvature adopt a non-lamellar structure called hexagonal II phase. Both of these structures have been suggested to contribute to the permeabilization of the mitochondrial outer membrane that occurs during apoptosis. PC, phosphatidylcholine; PG, phosphatidylglycerol; PS, phosphatidylserine; PI, phosphatidylinositol; PE, phosphatidylethanolamine; DAG, diacylglycerol.

Epand and colleagues have reported that the permeabilization and destabilization of membranes that are induced by tBid are instead potentiated by lipids that promote negative membrane curvature and a lamellar-to-hexagonal II phase transition (Fig. 4), and are inhibited by compounds, such as CsA, that favor positive curvature (Epand et al., 2002a). Formation of non-lamellar structures is known to be important for membrane fusion and fission events (Melikyan and Chernomordik, 1997; Ortiz et al., 1999) and has also been suggested to occur in the plasma membranes of cells undergoing apoptosis or necrosis (Aguilar et al., 1999). Interestingly, cardiolipin tends to adopt a hexagonal II phase in the presence of divalent cations that reduce the electrostatic repulsion between its negatively charged headgroups (Ortiz et al., 1999), and an increase in cytosolic Ca2+ levels is often observed during apoptosis. Epand and colleagues have also reported that permeabilization of liposomes by oligomeric BaxΔC requires Ca2+ and/or Mg2+ (Epand et al., 2002b; Epand et al., 2004). However, they also observed that it is activated by fatty acids and inhibited by LPC and dilysocardiolipin, even if all these lipids promote positive membrane curvature (Epand et al., 2004). Therefore, changing membrane curvature might not be the only mechanism through which lipids can influence the permeabilization of liposomes by activated Bax.

The results obtained in studies of synthetic liposomes must, however, be considered with caution, because the structure adopted by lipids is profoundly influenced by the experimental conditions, including salt concentration, pH, presence of divalent cations, temperature, transmembrane potential and nature of the fatty acyl chains of the phospholipids (Lewis and McElhaney, 2000; Ortiz et al., 1999). Moreover, artificial membranes do not reflect some of the important properties of biological membranes, in particular their mosaic nature, with organized protein and lipid microdomains. Understanding the importance of the lipidic composition of mitochondria for their permeabilization by Bax and Bak will certainly require working with cells and knocking down of enzymes involved in the biogenesis and the metabolism of mitochondrial lipids.

The number of molecules suggested to be involved in the regulation of apoptosis is increasing daily, yet important doubts remain concerning the mechanism of permeabilization of the mitochondrial outer membrane. The multiple binding partners reported for proteins of the Bcl-2 family could reflect the need to integrate information about the complex state of a cell. However, care must be taken, because Bcl-2 proteins are sticky and their conformation is influenced by the experimental conditions, in particular the presence of detergents. Most studies are carried out with cancer cell lines, in which components of the death pathways are altered. We must therefore emphasize the need to confirm observations in several cell types and ideally in primary cultures. Finally, even though lipids certainly play an important role in mitochondrial permeabilization, it remains very difficult to analyze their contributions in vivo. Therefore indirect evidence (e.g. analysis of Bax/Bak activation and cytochrome c release after changing the expression levels of lipid-modifying enzymes) will be needed, and innovative methods will have to be developed.

Apoptosis and necrosis are only extremes of an important array of mechanisms by which cells can die (Leist and Jaattela, 2001; Lockshin and Zakeri, 2002). Indeed, proteolytic degradation and autophagy can share some of the morphological and biochemical changes that characterize apoptosis. Therefore, some of the discrepancies in literature could be explained by the use of parallel but distinct programmed cell death mechanisms. Understanding the differences between these pathways will certainly prove crucial to the development of new therapeutic agents to provoke the death of cancer cells, in which one or several of these pathways may be inactivated.

We thank E.A.C. Lucken and J. Häsler for helpful comments and enthusiastic discussions about the manuscript.

Aguilar, L., Ortega-Pierres, G., Campos, B., Fonseca, R., Ibanez, M., Wong, C., Farfan, N., Naciff, J. M., Kaetzel, M. A., Dedman, J. R. et al. (
1999
). Phospholipid membranes form specific nonbilayer molecular arrangements that are antigenic.
J. Biol. Chem.
274
,
25193
-25196.
Andley, U. P., Song, Z., Wawrousek, E. F., Fleming, T. P. and Bassnett, S. (
2000
). Differential protective activity of alphaA- and alphaB-crystallin in lens epithelial cells.
J. Biol. Chem.
275
,
36823
-36831.
Antonsson, B., Montessuit, S., Lauper, S., Eskes, R. and Martinou, J. C. (
2000
). Bax oligomerization is required for channel-forming activity in liposomes and to trigger cytochrome c release from mitochondria.
Biochem. J.
345
,
271
-278.
Ardail, D., Privat, J. P., Egret-Charlier, M., Levrat, C., Lerme, F. and Louisot, P. (
1990
). Mitochondrial contact sites. Lipid composition and dynamics.
J. Biol. Chem.
265
,
18797
-18802.
Basañez, G., Nechushtan, A., Drozhinin, O., Chanturiya, A., Choe, E., Tutt, S., Wood, K. A., Hsu, Y., Zimmerberg, J. and Youle, R. J. (
1999
). Bax, but not Bcl-xL, decreases the lifetime of planar phospholipid bilayer membranes at subnanomolar concentrations.
Proc. Natl. Acad. Sci. USA
96
,
5492
-5497.
Basañez, G., Sharpe, J. C., Galanis, J., Brandt, T. B., Hardwick, J. M. and Zimmerberg, J. (
2002
). Bax-type apoptotic proteins porate pure lipid bilayers through a mechanism sensitive to intrinsic monolayer curvature.
J. Biol. Chem.
277
,
49360
-49365.
Borner, C. (
2003
). The Bcl-2 protein family: sensors and checkpoints for life- or-death decisions.
Mol. Immunol.
39
,
615
-647.
Brady, J. P., Garland, D., Duglas-Tabor, Y., Robison, W. G., Jr, Groome, A. and Wawrousek, E. F. (
1997
). Targeted disruption of the mouse alpha A-crystallin gene induces cataract and cytoplasmic inclusion bodies containing the small heat shock protein alpha B-crystallin.
Proc. Natl. Acad. Sci. USA
94
,
884
-889.
Brady, J. P., Garland, D. L., Green, D. E., Tamm, E. R., Giblin, F. J. and Wawrousek, E. F. (
2001
). AlphaB-crystallin in lens development and muscle integrity: a gene knockout approach.
Invest. Ophthalmol. Vis. Sci.
42
,
2924
-2934.
Calnan, B. J., Szychowski, S., Chan, F. K., Cado, D. and Winoto, A. (
1995
). A role for the orphan steroid receptor Nur77 in apoptosis accompanying antigen-induced negative selection.
Immunity
3
,
273
-282.
Cheng, E. H., Kirsch, D. G., Clem, R. J., Ravi, R., Kastan, M. B., Bedi, A., Ueno, K. and Hardwick, J. M. (
1997
). Conversion of Bcl-2 to a Bax-like death effector by caspases.
Science
278
,
1966
-1968.
Cheng, E. H., Wei, M. C., Weiler, S., Flavell, R. A., Mak, T. W., Lindsten, T. and Korsmeyer, S. J. (
2001
). BCL-2, BCL-X(L) sequester BH3 domain-only molecules preventing BAX- and BAK-mediated mitochondrial apoptosis.
Mol. Cell
8
,
705
-711.
Cheng, E. H., Sheiko, T. V., Fisher, J. K., Craigen, W. J. and Korsmeyer, S. J. (
2003
). VDAC2 inhibits BAK activation and mitochondrial apoptosis.
Science
301
,
513
-517.
Chipuk, J. E. and Green, D. R. (
2003
). p53's believe it or not: lessons on transcription-independent death.
J. Clin. Immunol.
23
,
355
-361.
Chipuk, J. E., Kuwana, T., Bouchier-Hayes, L., Droin, N. M., Newmeyer, D. D., Schuler, M. and Green, D. R. (
2004
). Direct activation of Bax by p53 mediates mitochondrial membrane permeabilization and apoptosis.
Science
303
,
1010
-1014.
Chua, B. T., Volbracht, C., Tan, K. O., Li, R., Yu, V. C. and Li, P. (
2003
). Mitochondrial translocation of cofilin is an early step in apoptosis induction.
Nat. Cell Biol.
5
,
1083
-1089.
Clem, R. J., Cheng, E. H., Karp, C. L., Kirsch, D. G., Ueno, K., Takahashi, A., Kastan, M. B., Griffin, D. E., Earnshaw, W. C., Veliuona, M. A. et al. (
1998
). Modulation of cell death by Bcl-XL through caspase interaction.
Proc. Natl. Acad. Sci. USA
95
,
554
-559.
Cory, S., Huang, D. C. and Adams, J. M. (
2003
). The Bcl-2 family: roles in cell survival and oncogenesis.
Oncogene
22
,
8590
-8607.
Cristea, I. M. and Degli Esposti, M. (
2004
). Membrane lipids and cell death: an overview.
Chem. Phys. Lipids
129
,
133
-160.
Crompton, M., Virji, S. and Ward, J. M. (
1998
). Cyclophilin-D binds strongly to complexes of the voltage-dependent anion channel and the adenine nucleotide translocase to form the permeability transition pore.
Eur. J. Biochem.
258
,
729
-735.
Crompton, M., Barksby, E., Johnson, N. and Capano, M. (
2002
). Mitochondrial intermembrane junctional complexes and their involvement in cell death.
Biochimie
84
,
143
-152.
Cuddeback, S. M., Yamaguchi, H., Komatsu, K., Miyashita, T., Yamada, M., Wu, C., Singh, S. and Wang, H. G. (
2001
). Molecular cloning and characterization of Bif-1. A novel Src homology 3 domain-containing protein that associates with Bax.
J. Biol. Chem.
276
,
20559
-20565.
Degli Esposti, M. (
2002
). Sequence and functional similarities between pro-apoptotic Bid and plant lipid transfer proteins.
Biochim. Biophys. Acta
1553
,
331
-340.
Desagher, S. and Martinou, J. C. (
2000
). Mitochondria as the central control point of apoptosis.
Trends Cell Biol.
10
,
369
-377.
Desagher, S., Osen-Sand, A., Nichols, A., Eskes, R., Montessuit, S., Lauper, S., Maundrell, K., Antonsson, B. and Martinou, J. C. (
1999
). Bid-induced conformational change of Bax is responsible for mitochondrial cytochrome c release during apoptosis.
J. Cell Biol.
144
,
891
-901.
Ding, H. F., McGill, G., Rowan, S., Schmaltz, C., Shimamura, A. and Fisher, D. E. (
1998
). Oncogene-dependent regulation of caspase activation by p53 protein in a cell-free system.
J. Biol. Chem.
273
,
28378
-28383.
Distelhorst, C. W. and Shore, G. C. (
2004
). Bcl-2 and calcium: controversy beneath the surface.
Oncogene
23
,
2875
-2880.
Dougherty, M. K. and Morrison, D. K. (
2004
). Unlocking the code of 14-3-3.
J. Cell Sci.
117
,
1875
-1884.
Epand, R. M., Epand, R. F. and McKenzie, R. C. (
1987
). Effects of viral chemotherapeutic agents on membrane properties. Studies of cyclosporin A, benzyloxycarbonyl-D-Phe-L-Phe-Gly and amantadine.
J. Biol. Chem.
262
,
1526
-1529.
Epand, R. F., Martinou, J. C., Fornallaz-Mulhauser, M., Hughes, D. W. and Epand, R. M. (
2002a
). The apoptotic protein tBid promotes leakage by altering membrane curvature.
J. Biol. Chem.
277
,
32632
-32639.
Epand, R. F., Martinou, J. C., Montessuit, S. and Epand, R. M. (
2002b
). Membrane perturbations induced by the apoptotic Bax protein.
Biochem. J.
367
,
849
-855.
Epand, R. F., Martinou, J. C., Montessuit, S., Epand, R. M. and Yip, C. M. (
2002c
). Direct evidence for membrane pore formation by the apoptotic protein Bax.
Biochem. Biophys. Res. Commun.
298
,
744
-749.
Epand, R. F., Martinou, J. C., Montessuit, S. and Epand, R. M. (
2003
). Transbilayer lipid diffusion promoted by Bax: implications for apoptosis.
Biochemistry
42
,
14576
-14582.
Epand, R. F., Martinou, J. C., Montessuit, S. and Epand, R. M. (
2004
). Fatty acids enhance membrane permeabilization by pro-apoptotic Bax.
Biochem. J.
377
,
509
-516.
Erster, S., Mihara, M., Kim, R. H., Petrenko, O. and Moll, U. M. (
2004
). In vivo mitochondrial p53 translocation triggers a rapid first wave of cell death in response to DNA damage that can precede p53 target gene activation.
Mol. Cell. Biol.
24
,
6728
-6741.
Eskes, R., Desagher, S., Antonsson, B. and Martinou, J. C. (
2000
). Bid induces the oligomerization and insertion of Bax into the outer mitochondrial membrane.
Mol. Cell. Biol.
20
,
929
-935.
Esposti, M. D., Erler, J. T., Hickman, J. A. and Dive, C. (
2001
). Bid, a widely expressed proapoptotic protein of the Bcl-2 family, displays lipid transfer activity.
Mol. Cell. Biol.
21
,
7268
-7276.
Esposti, M. D., Cristea, I. M., Gaskell, S. J., Nakao, Y. and Dive, C. (
2003
). Proapoptotic Bid binds to monolysocardiolipin, a new molecular connection between mitochondrial membranes and cell death.
Cell Death Differ.
10
,
1300
-1309.
Fernandez-Salas, E., Sagar, M., Cheng, C., Yuspa, S. H. and Weinberg, W. C. (
1999
). p53 and tumor necrosis factor alpha regulate the expression of a mitochondrial chloride channel protein.
J. Biol. Chem.
274
,
36488
-36497.
Fernandez-Salas, E., Suh, K. S., Speransky, V. V., Bowers, W. L., Levy, J. M., Adams, T., Pathak, K. R., Edwards, L. E., Hayes, D. D., Cheng, C. et al. (
2002
). mtCLIC/CLIC4, an organellular chloride channel protein, is increased by DNA damage and participates in the apoptotic response to p53.
Mol. Cell. Biol.
22
,
3610
-3620.
Fischer, U., Janicke, R. U. and Schulze-Osthoff, K. (
2003
). Many cuts to ruin: a comprehensive update of caspase substrates.
Cell Death Differ.
10
,
76
-100.
Frank, S., Gaume, B., Bergmann-Leitner, E. S., Leitner, W. W., Robert, E. G., Catez, F., Smith, C. L. and Youle, R. J. (
2001
). The role of dynamin-related protein 1, a mediator of mitochondrial fission, in apoptosis.
Dev. Cell
1
,
515
-525.
Fruman, D. A., Klee, C. B., Bierer, B. E. and Burakoff, S. J. (
1992
). Calcineurin phosphatase activity in T lymphocytes is inhibited by FK 506 and cyclosporin A.
Proc. Natl. Acad. Sci. USA
89
,
3686
-3690.
Garcia Fernandez, M., Troiano, L., Moretti, L., Nasi, M., Pinti, M., Salvioli, S., Dobrucki, J. and Cossarizza, A. (
2002
). Early changes in intramitochondrial cardiolipin distribution during apoptosis.
Cell Growth Differ.
13
,
449
-455.
Goping, I. S., Gross, A., Lavoie, J. N., Nguyen, M., Jemmerson, R., Roth, K., Korsmeyer, S. J. and Shore, G. C. (
1998
). Regulated targeting of BAX to mitochondria.
J. Cell Biol.
143
,
207
-215.
Gotoh, T., Terada, K., Oyadomari, S. and Mori, M. (
2004
). hsp70-DnaJ chaperone pair prevents nitric oxide- and CHOP-induced apoptosis by inhibiting translocation of Bax to mitochondria.
Cell Death Differ.
11
,
390
-402.
Gross, A., McDonnell, J. M. and Korsmeyer, S. J. (
1999
). BCL-2 family members and the mitochondria in apoptosis.
Genes Dev.
13
,
1899
-1911.
Gu, Y., Seidl, K. J., Rathbun, G. A., Zhu, C., Manis, J. P., van der Stoep, N., Davidson, L., Cheng, H. L., Sekiguchi, J. M., Frank, K. et al. (
1997
). Growth retardation and leaky SCID phenotype of Ku70-deficient mice.
Immunity
7
,
653
-665.
Guo, B., Zhai, D., Cabezas, E., Welsh, K., Nouraini, S., Satterthwait, A. C. and Reed, J. C. (
2003
). Humanin peptide suppresses apoptosis by interfering with Bax activation.
Nature
423
,
456
-461.
Gustafsson, A. B., Tsai, J. G., Logue, S. E., Crow, M. T. and Gottlieb, R. A. (
2004
). Apoptosis repressor with caspase recruitment domain protects against cell death by interfering with Bax activation.
J. Biol. Chem.
279
,
21233
-21238.
Halestrap, A. P., McStay, G. P. and Clarke, S. J. (
2002
). The permeability transition pore complex: another view.
Biochimie
84
,
153
-166.
Hashimoto, Y., Niikura, T., Tajima, H., Yasukawa, T., Sudo, H., Ito, Y., Kita, Y., Kawasumi, M., Kouyama, K., Doyu, M. et al. (
2001
). A rescue factor abolishing neuronal cell death by a wide spectrum of familial Alzheimer's disease genes and Abeta.
Proc. Natl. Acad. Sci. USA
98
,
6336
-6341.
Haupt, S., Berger, M., Goldberg, Z. and Haupt, Y. (
2003
). Apoptosis - the p53 network.
J. Cell Sci.
116
,
4077
-4085.
He, L., Perkins, G. A., Poblenz, A. T., Harris, J. B., Hung, M., Ellisman, M. H. and Fox, D. A. (
2003
). Bcl-xL overexpression blocks bax-mediated mitochondrial contact site formation and apoptosis in rod photoreceptors of lead-exposed mice.
Proc. Natl. Acad. Sci. USA
100
,
1022
-1027.
Hitomi, J., Katayama, T., Eguchi, Y., Kudo, T., Taniguchi, M., Koyama, Y., Manabe, T., Yamagishi, S., Bando, Y., Imaizumi, K. et al. (
2004
). Involvement of caspase-4 in endoplasmic reticulum stress-induced apoptosis and Abeta-induced cell death.
J. Cell Biol.
165
,
347
-356.
Ho, J. and Benchimol, S. (
2003
). Transcriptional repression mediated by the p53 tumour suppressor.
Cell Death Differ.
10
,
404
-408.
Hovius, R., Thijssen, J., van der Linden, P., Nicolay, K. and de Kruijff, B. (
1993
). Phospholipid asymmetry of the outer membrane of rat liver mitochondria. Evidence for the presence of cardiolipin on the outside of the outer membrane.
FEBS Lett.
330
,
71
-76.
Hsu, Y. T. and Youle, R. J. (
1997
). Nonionic detergents induce dimerization among members of the Bcl-2 family.
J. Biol. Chem.
272
,
13829
-13834.
Hsu, S. Y., Kaipia, A., McGee, E., Lomeli, M. and Hsueh, A. J. (
1997a
). Bok is a pro-apoptotic Bcl-2 protein with restricted expression in reproductive tissues and heterodimerizes with selective anti-apoptotic Bcl-2 family members.
Proc. Natl. Acad. Sci. USA
94
,
12401
-12406.
Hsu, Y. T., Wolter, K. G. and Youle, R. J. (
1997b
). Cytosol-to-membrane redistribution of Bax and Bcl-X(L) during apoptosis.
Proc. Natl. Acad. Sci. USA
94
,
3668
-3672.
Hunt, C. R., Dix, D. J., Sharma, G. G., Pandita, R. K., Gupta, A., Funk, M. and Pandita, T. K. (
2004
). Genomic instability and enhanced radiosensitivity in Hsp70.1- and Hsp70.3-deficient mice.
Mol. Cell. Biol.
24
,
899
-911.
Huttner, W. B. and Schmidt, A. A. (
2002
). Membrane curvature: a case of endofeelin'.
Trends Cell Biol.
12
,
155
-158.
Inohara, N., Ekhterae, D., Garcia, I., Carrio, R., Merino, J., Merry, A., Chen, S. and Nunez, G. (
1998
). Mtd, a novel Bcl-2 family member activates apoptosis in the absence of heterodimerization with Bcl-2 and Bcl-XL.
J. Biol. Chem.
273
,
8705
-8710.
Israelachvili, J. N. and Mitchell, D. J. (
1975
). A model for the packing of lipids in bilayer membranes.
Biochim. Biophys. Acta
389
,
13
-19.
Jang, J. H. and Surh, Y. J. (
2003
). Potentiation of cellular antioxidant capacity by Bcl-2: implications for its antiapoptotic function.
Biochem. Pharmacol.
66
,
1371
-1379.
Kamradt, M. C., Chen, F., Sam, S. and Cryns, V. L. (
2002
). The small heat shock protein alpha B-crystallin negatively regulates apoptosis during myogenic differentiation by inhibiting caspase-3 activation.
J. Biol. Chem.
277
,
38731
-38736.
Karbowski, M. and Youle, R. J. (
2003
). Dynamics of mitochondrial morphology in healthy cells and during apoptosis.
Cell Death Differ.
10
,
870
-880.
Karbowski, M., Jeong, S. Y. and Youle, R. J. (
2004
). Endophilin B1 is required for the maintenance of mitochondrial morphology.
J. Cell Biol.
166
,
1027
-1039.
Kim, P. K., Annis, M. G., Dlugosz, P. J., Leber, B. and Andrews, D. W. (
2004a
). During apoptosis bcl-2 changes membrane topology at both the endoplasmic reticulum and mitochondria.
Mol. Cell
14
,
523
-529.
Kim, T. H., Zhao, Y., Ding, W. X., Shin, J. N., He, X., Seo, Y. W., Chen, J., Rabinowich, H., Amoscato, A. A. and Yin, X. M. (
2004b
). Bid-cardiolipin interaction at mitochondrial contact site contributes to mitochondrial cristae reorganization and cytochrome C release.
Mol. Biol. Cell
15
,
3061
-3072.
Kokoszka, J. E., Waymire, K. G., Levy, S. E., Sligh, J. E., Cai, J., Jones, D. P., MacGregor, G. R. and Wallace, D. C. (
2004
). The ADP/ATP translocator is not essential for the mitochondrial permeability transition pore.
Nature
427
,
461
-465.
Konishi, A., Shimizu, S., Hirota, J., Takao, T., Fan, Y., Matsuoka, Y., Zhang, L., Yoneda, Y., Fujii, Y., Skoultchi, A. I. et al. (
2003
). Involvement of histone H1.2 in apoptosis induced by DNA double-strand breaks.
Cell
114
,
673
-688.
Koseki, T., Inohara, N., Chen, S. and Nunez, G. (
1998
). ARC, an inhibitor of apoptosis expressed in skeletal muscle and heart that interacts selectively with caspases.
Proc. Natl. Acad. Sci. USA
95
,
5156
-5160.
Kuwana, T., Mackey, M. R., Perkins, G., Ellisman, M. H., Latterich, M., Schneiter, R., Green, D. R. and Newmeyer, D. D. (
2002
). Bid, Bax, and lipids cooperate to form supramolecular openings in the outer mitochondrial membrane.
Cell
111
,
331
-342.
Lassus, P., Opitz-Araya, X. and Lazebnik, Y. (
2002
). Requirement for caspase-2 in stress-induced apoptosis before mitochondrial permeabilization.
Science
297
,
1352
-1354.
Lee, S. L., Wesselschmidt, R. L., Linette, G. P., Kanagawa, O., Russell, J. H. and Milbrandt, J. (
1995
). Unimpaired thymic and peripheral T cell death in mice lacking the nuclear receptor NGFI-B (Nur77).
Science
269
,
532
-535.
Leist, M. and Jaattela, M. (
2001
). Four deaths and a funeral: from caspases to alternative mechanisms.
Nat. Rev. Mol. Cell Biol.
2
,
589
-598.
Leu, J. I., Dumont, P., Hafey, M., Murphy, M. E. and George, D. L. (
2004
). Mitochondrial p53 activates Bak and causes disruption of a Bak-Mcl1 complex.
Nat. Cell Biol.
6
,
443
-450.
Lewis, R. N. and McElhaney, R. N. (
2000
). Surface charge markedly attenuates the nonlamellar phase-forming propensities of lipid bilayer membranes: calorimetric and (31)P-nuclear magnetic resonance studies of mixtures of cationic, anionic, and zwitterionic lipids.
Biophys. J.
79
,
1455
-1464.
Li, P. F., Dietz, R. and von Harsdorf, R. (
1999
). p53 regulates mitochondrial membrane potential through reactive oxygen species and induces cytochrome c-independent apoptosis blocked by Bcl-2.
EMBO J.
18
,
6027
-6036.
Li, H., Kolluri, S. K., Gu, J., Dawson, M. I., Cao, X., Hobbs, P. D., Lin, B., Chen, G., Lu, J., Lin, F. et al. (
2000
). Cytochrome c release and apoptosis induced by mitochondrial targeting of nuclear orphan receptor TR3.
Science
289
,
1159
-1164.
Lin, Y., Ma, W. and Benchimol, S. (
2000
). Pidd, a new death-domain-containing protein, is induced by p53 and promotes apoptosis.
Nat. Genet.
26
,
122
-127.
Lin, B., Kolluri, S. K., Lin, F., Liu, W., Han, Y. H., Cao, X., Dawson, M. I., Reed, J. C. and Zhang, X. K. (
2004
). Conversion of Bcl-2 from protector to killer by interaction with nuclear orphan receptor Nur77/TR3.
Cell
116
,
527
-540.
Lockshin, R. A. and Zakeri, Z. (
2002
). Caspase-independent cell deaths.
Curr. Opin. Cell Biol.
14
,
727
-733.
Lutter, M., Fang, M., Luo, X., Nishijima, M., Xie, X. and Wang, X. (
2000
). Cardiolipin provides specificity for targeting of tBid to mitochondria.
Nat. Cell Biol.
2
,
754
-761.
Majewski, N., Nogueira, V., Robey, R. B. and Hay, N. (
2004
). Akt inhibits apoptosis downstream of BID cleavage via a glucose-dependent mechanism involving mitochondrial hexokinases.
Mol. Cell. Biol.
24
,
730
-740.
Makin, G. W., Corfe, B. M., Griffiths, G. J., Thistlethwaite, A., Hickman, J. A. and Dive, C. (
2001
). Damage-induced Bax N-terminal change, translocation to mitochondria and formation of Bax dimers/complexes occur regardless of cell fate.
EMBO J.
20
,
6306
-6315.
Mao, Y. W., Liu, J. P., Xiang, H. and Li, D. W. (
2004
). Human alphaA- and alphaB-crystallins bind to Bax and Bcl-X(S) to sequester their translocation during staurosporine-induced apoptosis.
Cell Death Differ.
11
,
512
-526.
Marani, M., Tenev, T., Hancock, D., Downward, J. and Lemoine, N. R. (
2002
). Identification of novel isoforms of the BH3 domain protein Bim which directly activate Bax to trigger apoptosis.
Mol. Cell. Biol.
22
,
3577
-3589.
Marchenko, N. D., Zaika, A. and Moll, U. M. (
2000
). Death signal-induced localization of p53 protein to mitochondria. A potential role in apoptotic signaling.
J. Biol. Chem.
275
,
16202
-16212.
Mariathasan, S., Newton, K., Monack, D. M., Vucic, D., French, D. M., Lee, W. P., Roose-Girma, M., Erickson, S. and Dixit, V. M. (
2004
). Differential activation of the inflammasome by caspase-1 adaptors ASC and Ipaf.
Nature
430
,
213
-218.
Martinou, J. C. and Green, D. R. (
2001
). Breaking the mitochondrial barrier.
Nat. Rev. Mol. Cell Biol.
2
,
63
-67.
Matsko, C. M., Hunter, O. C., Rabinowich, H., Lotze, M. T. and Amoscato, A. A. (
2001
). Mitochondrial lipid alterations during Fas- and radiation-induced apoptosis.
Biochem. Biophys. Res. Commun.
287
,
1112
-1120.
Matsuda, K., Yoshida, K., Taya, Y., Nakamura, K., Nakamura, Y. and Arakawa, H. (
2002
). p53AIP1 regulates the mitochondrial apoptotic pathway.
Cancer Res.
62
,
2883
-2889.
Matsuzaki, K., Sugishita, K., Ishibe, N., Ueha, M., Nakata, S., Miyajima, K. and Epand, R. M. (
1998
). Relationship of membrane curvature to the formation of pores by magainin 2.
Biochemistry
37
,
11856
-11863.
McGuinness, O., Yafei, N., Costi, A. and Crompton, M. (
1990
). The presence of two classes of high-affinity cyclosporin A binding sites in mitochondria. Evidence that the minor component is involved in the opening of an inner-membrane Ca(2+)-dependent pore.
Eur. J. Biochem.
194
,
671
-679.
McMillin, J. B. and Dowhan, W. (
2002
). Cardiolipin and apoptosis.
Biochim. Biophys. Acta
1585
,
97
-107.
Melikyan, G. B. and Chernomordik, L. V. (
1997
). Membrane rearrangements in fusion mediated by viral proteins.
Trends Microbiol.
5
,
349
-355.
Mihara, M., Erster, S., Zaika, A., Petrenko, O., Chittenden, T., Pancoska, P. and Moll, U. M. (
2003
). p53 has a direct apoptogenic role at the mitochondria.
Mol. Cell
11
,
577
-590.
Modregger, J., Schmidt, A. A., Ritter, B., Huttner, W. B. and Plomann, M. (
2003
). Characterization of Endophilin B1b, a brain-specific membrane-associated lysophosphatidic acid acyl transferase with properties distinct from endophilin A1.
J. Biol. Chem.
278
,
4160
-4167.
Morishima, N., Nakanishi, K., Takenouchi, H., Shibata, T. and Yasuhiko, Y. (
2002
). An endoplasmic reticulum stress-specific caspase cascade in apoptosis. Cytochrome c-independent activation of caspase-9 by caspase-12.
J. Biol. Chem.
277
,
34287
-34294.
Moss, M. L., Palmer, R. E., Kuzmic, P., Dunlap, B. E., Henzel, W., Kofron, J. L., Mellon, W. S., Royer, C. A. and Rich, D. H. (
1992
). Identification of actin and HSP 70 as cyclosporin A binding proteins by photoaffinity labeling and fluorescence displacement assays.
J. Biol. Chem.
267
,
22054
-22059.
Nakagawa, T., Zhu, H., Morishima, N., Li, E., Xu, J., Yankner, B. A. and Yuan, J. (
2000
). Caspase-12 mediates endoplasmic-reticulum-specific apoptosis and cytotoxicity by amyloid-beta.
Nature
403
,
98
-103.
Nam, Y. J., Mani, K., Ashton, A. W., Peng, C. F., Krishnamurthy, B., Hayakawa, Y., Lee, P., Korsmeyer, S. J. and Kitsis, R. N. (
2004
). Inhibition of both the extrinsic and intrinsic death pathways through nonhomotypic death-fold interactions.
Mol. Cell
15
,
901
-912.
Nechushtan, A., Smith, C. L., Hsu, Y. T. and Youle, R. J. (
1999
). Conformation of the Bax C-terminus regulates subcellular location and cell death.
EMBO J.
18
,
2330
-2341.
Nicholson, D. W. (
1999
). Caspase structure, proteolytic substrates, and function during apoptotic cell death.
Cell Death Differ.
6
,
1028
-1042.
Nomura, K., Imai, H., Koumura, T., Kobayashi, T. and Nakagawa, Y. (
2000
). Mitochondrial phospholipid hydroperoxide glutathione peroxidase inhibits the release of cytochrome c from mitochondria by suppressing the peroxidation of cardiolipin in hypoglycaemia-induced apoptosis.
Biochem. J.
351
,
183
-193.
Nomura, M., Shimizu, S., Sugiyama, T., Narita, M., Ito, T., Matsuda, H. and Tsujimoto, Y. (
2003
). 14-3-3 interacts directly with and negatively regulates pro-apoptotic Bax.
J. Biol. Chem.
278
,
2058
-2065.
Oda, K., Arakawa, H., Tanaka, T., Matsuda, K., Tanikawa, C., Mori, T., Nishimori, H., Tamai, K., Tokino, T., Nakamura, Y. et al. (
2000
). p53AIP1, a potential mediator of p53-dependent apoptosis, and its regulation by Ser-46-phosphorylated p53.
Cell
102
,
849
-862.
Ohtsuka, T., Ryu, H., Minamishima, Y. A., Macip, S., Sagara, J., Nakayama, K. I., Aaronson, S. A. and Lee, S. W. (
2004
). ASC is a Bax adaptor and regulates the p53-Bax mitochondrial apoptosis pathway.
Nat. Cell Biol.
6
,
121
-128.
Ortiz, A., Killian, J. A., Verkleij, A. J. and Wilschut, J. (
1999
). Membrane fusion and the lamellar-to-inverted-hexagonal phase transition in cardiolipin vesicle systems induced by divalent cations.
Biophys. J.
77
,
2003
-2014.
Ostrander, D. B., Sparagna, G. C., Amoscato, A. A., McMillin, J. B. and Dowhan, W. (
2001
). Decreased cardiolipin synthesis corresponds with cytochrome c release in palmitate-induced cardiomyocyte apoptosis.
J. Biol. Chem.
276
,
38061
-38067.
Ott, M., Robertson, J. D., Gogvadze, V., Zhivotovsky, B. and Orrenius, S. (
2002
). Cytochrome c release from mitochondria proceeds by a two-step process.
Proc. Natl. Acad. Sci. USA
99
,
1259
-1263.
Pastorino, J. G., Shulga, N. and Hoek, J. B. (
2002
). Mitochondrial binding of hexokinase II inhibits Bax-induced cytochrome c release and apoptosis.
J. Biol. Chem.
277
,
7610
-7618.
Perez, D. and White, E. (
2000
). TNF-alpha signals apoptosis through a bid-dependent conformational change in Bax that is inhibited by E1B 19K.
Mol. Cell
6
,
53
-63.
Petros, A. M., Olejniczak, E. T. and Fesik, S. W. (
2004
). Structural biology of the Bcl-2 family of proteins.
Biochim. Biophys. Acta
1644
,
83
-94.
Petrosillo, G., Ruggiero, F. M., Pistolese, M. and Paradies, G. (
2001
). Reactive oxygen species generated from the mitochondrial electron transport chain induce cytochrome c dissociation from beef-heart submitochondrial particles via cardiolipin peroxidation. Possible role in the apoptosis.
FEBS Lett.
509
,
435
-438.
Puthalakath, H. and Strasser, A. (
2002
). Keeping killers on a tight leash: transcriptional and post-translational control of the pro-apoptotic activity of BH3-only proteins.
Cell Death Differ.
9
,
505
-512.
Qi, L., Danielson, N. D., Dai, Q. and Lee, R. M. (
2003
). Capillary electrophoresis of cardiolipin with on-line dye interaction and spectrophotometric detection.
Electrophoresis
24
,
1680
-1686.
Ranger, A. M., Malynn, B. A. and Korsmeyer, S. J. (
2001
). Mouse models of cell death.
Nat. Genet.
28
,
113
-118.
Roucou, X., Montessuit, S., Antonsson, B. and Martinou, J. C. (
2002
). Bax oligomerization in mitochondrial membranes requires tBid (caspase-8-cleaved Bid) and a mitochondrial protein.
Biochem. J.
368
,
915
-921.
Ruffolo, S. C. and Shore, G. C. (
2003
). BCL-2 selectively interacts with the BID-induced open conformer of BAK, inhibiting BAK auto-oligomerization.
J. Biol. Chem.
278
,
25039
-25045.
Saito, M., Korsmeyer, S. J. and Schlesinger, P. H. (
2000
). BAX-dependent transport of cytochrome c reconstituted in pure liposomes.
Nat. Cell Biol.
2
,
553
-555.
Samuel, T., Weber, H. O., Rauch, P., Verdoodt, B., Eppel, J. T., McShea, A., Hermeking, H. and Funk, J. O. (
2001
). The G2/M regulator 14-3-3sigma prevents apoptosis through sequestration of Bax.
J. Biol. Chem.
276
,
45201
-45206.
Sawada, M., Sun, W., Hayes, P., Leskov, K., Boothman, D. A. and Matsuyama, S. (
2003
). Ku70 suppresses the apoptotic translocation of Bax to mitochondria.
Nat. Cell Biol.
5
,
320
-329.
Scaffidi, C., Fulda, S., Srinivasan, A., Friesen, C., Li, F., Tomaselli, K. J., Debatin, K. M., Krammer, P. H. and Peter, M. E. (
1998
). Two CD95 (APO-1/Fas) signaling pathways.
EMBO J.
17
,
1675
-1687.
Schinzel, A., Kaufmann, T. and Borner, C. (
2004a
). Bcl-2 family members: integrators of survival and death signals in physiology and pathology.
Biochim. Biophys. Acta
1644
,
95
-105.
Schinzel, A., Kaufmann, T., Schuler, M., Martinalbo, J., Grubb, D. and Borner, C. (
2004b
). Conformational control of Bax localization and apoptotic activity by Pro168.
J. Cell Biol.
164
,
1021
-1032.
Scorrano, L., Ashiya, M., Buttle, K., Weiler, S., Oakes, S. A., Mannella, C. A. and Korsmeyer, S. J. (
2002
). A distinct pathway remodels mitochondrial cristae and mobilizes cytochrome c during apoptosis.
Dev. Cell
2
,
55
-67.
Shidoji, Y., Hayashi, K., Komura, S., Ohishi, N. and Yagi, K. (
1999
). Loss of molecular interaction between cytochrome c and cardiolipin due to lipid peroxidation.
Biochem. Biophys. Res. Commun.
264
,
343
-347.
Slee, E. A., O'Connor, D. J. and Lu, X. (
2004
). To die or not to die: how does p53 decide?
Oncogene
23
,
2809
-2818.
Sorice, M., Circella, A., Cristea, I. M., Garofalo, T., Renzo, L. D., Alessandri, C., Valesini, G. and Esposti, M. D. (
2004
). Cardiolipin and its metabolites move from mitochondria to other cellular membranes during death receptor-mediated apoptosis.
Cell Death Differ.
11
,
1133
-1145.
Suh, K. S., Mutoh, M., Nagashima, K., Fernandez-Salas, E., Edwards, L. E., Hayes, D. D., Crutchley, J. M., Marin, K. G., Dumont, R. A., Levy, J. M. et al. (
2004
). The organellular chloride channel protein CLIC4/mtCLIC translocates to the nucleus in response to cellular stress and accelerates apoptosis.
J. Biol. Chem.
279
,
4632
-4641.
Suzuki, M., Youle, R. J. and Tjandra, N. (
2000
). Structure of Bax: coregulation of dimer formation and intracellular localization.
Cell
103
,
645
-654.
Tan, K. O., Tan, K. M., Chan, S. L., Yee, K. S., Bevort, M., Ang, K. C. and Yu, V. C. (
2001
). MAP-1, a novel proapoptotic protein containing a BH3-like motif that associates with Bax through its Bcl-2 homology domains.
J. Biol. Chem.
276
,
2802
-2807.
Terrones, O., Antonsson, B., Yamaguchi, H., Wang, H. G., Liu, Y., Lee, R. M., Herrmann, A. and Basañez, G. (
2004
). Lipidic pore formation by the concerted action of pro-apoptotic BAX and tBID.
J. Biol. Chem.
279
,
30081
-30091.
Tinel, A. and Tschopp, J. (
2004
). The PIDDosome, a protein complex implicated in activation of caspase-2 in response to genotoxic stress.
Science
304
,
843
-846.
Tsujimoto, Y. and Shimizu, S. (
2000
). VDAC regulation by the Bcl-2 family of proteins.
Cell Death Differ.
7
,
1174
-1181.
Van Loo, G., Demol, H., van Gurp, M., Hoorelbeke, B., Schotte, P., Beyaert, R., Zhivotovsky, B., Gevaert, K., Declercq, W., Vandekerckhove, J. et al. (
2002
). A matrix-assisted laser desorption ionization post-source decay (MALDI-PSD) analysis of proteins released from isolated liver mitochondria treated with recombinant truncated Bid.
Cell Death Differ.
9
,
301
-308.
Waterhouse, N. J., Ricci, J. E. and Green, D. R. (
2002
). And all of a sudden it's over: mitochondrial outer-membrane permeabilization in apoptosis.
Biochimie
84
,
113
-121.
Wei, M. C., Lindsten, T., Mootha, V. K., Weiler, S., Gross, A., Ashiya, M., Thompson, C. B. and Korsmeyer, S. J. (
2000
). tBID, a membrane-targeted death ligand, oligomerizes BAK to release cytochrome c.
Genes Dev.
14
,
2060
-2071.
Wei, M. C., Zong, W. X., Cheng, E. H., Lindsten, T., Panoutsakopoulou, V., Ross, A. J., Roth, K. A., MacGregor, G. R., Thompson, C. B. and Korsmeyer, S. J. (
2001
). Proapoptotic BAX and BAK: a requisite gateway to mitochondrial dysfunction and death.
Science
292
,
727
-730.
Wieckowski, M. R., Vyssokikh, M., Dymkowska, D., Antonsson, B., Brdiczka, D. and Wojtczak, L. (
2001
). Oligomeric C-terminal truncated Bax preferentially releases cytochrome c but not adenylate kinase from mitochondria, outer membrane vesicles and proteoliposomes.
FEBS Lett.
505
,
453
-459.
Willis, S., Day, C. L., Hinds, M. G. and Huang, D. C. (
2003
). The Bcl-2-regulated apoptotic pathway.
J. Cell Sci.
116
,
4053
-4056.
Wright, M. M., Howe, A. G. and Zaremberg, V. (
2004
). Cell membranes and apoptosis: role of cardiolipin, phosphatidylcholine, and anticancer lipid analogues.
Biochem. Cell Biol.
82
,
18
-26.
Zamzami, N. and Kroemer, G. (
2001
). The mitochondrion in apoptosis: how Pandora's box opens.
Nat. Rev. Mol. Cell Biol.
2
,
67
-71.
Zimmermann, K. C., Bonzon, C. and Green, D. R. (
2001
). The machinery of programmed cell death.
Pharmacol. Ther.
92
,
57
-70.