BCL-2 was the first antideath gene dis-covered, a milestone that effectively launched a new era in cell death research. Since its discovery more than 2 decades ago, multiple members of the human Bcl-2 family of apoptosis-regulating proteins have been identified, including 6 antiapoptotic proteins, 3 structurally similar proapoptotic proteins, and several structurally diverse proapoptotic interacting proteins that operate as upstream agonists or antagonists. Bcl-2–family proteins regulate all major types of cell death, including apoptosis, necrosis, and autophagy. As such, they operate as nodal points at the convergence of multiple pathways with broad relevance to biology and medicine. Bcl-2 derives its name from its original discovery in the context of B-cell lymphomas, where chromosomal translocations commonly activate the BCL-2 protooncogene, endowing B cells with a selective survival advantage that promotes their neoplastic expansion. The concept that defective programmed cell death contributes to malignancy was established by studies of Bcl-2, representing a major step forward in current understanding of tumorigenesis. Experimental therapies targeting Bcl-2 family mRNAs or proteins are currently in clinical testing, raising hopes that a new class of anticancer drugs may be near.

Cell death can be either physiologic or pathologic. Although medicine focused on pathologic cell death for centuries, it was the discovery of programmed cell death that breathed new life into the field of cell death research, as well as sparking major advances in multiple areas of physiology and medicine. Physiologic cell death in animal species generally occurs through a mechanism commonly called “apoptosis,” typically involving activation of intracellular proteases known as caspases.1  The proteolytic events mediated by caspases impart characteristic morphologic and ultrastructural changes to dying cells that define the apoptotic phenotype. Among the features of apoptosis are cell shrinkage, blebbing of the plasma membrane without loss of integrity, nuclear fragmentation, and chromatin condensation. In vivo, apoptotic cells are cleared by phagocytosis before they can rupture, maintaining ATP and ion homeostasis even as they are cleared from the body.

It is estimated that the average adult human produces and in parallel eradicates approximately 60 billion cells daily, with new cells formed by cell division and old cells eliminated by apoptosis, thus striking a balance under normal circumstances. This ability to control cell numbers at both the points of entry and exit allows our bodies to more rapidly respond to stress, such as mounting a white blood cell count in the face of acute infection. In this regard, hematopoietic growth factors (eg, granulocyte macrophage–colony-stimulating factor, granulocyte cell-stimulating factor, interleukin-3) typically transduce signals both for expansion and differentiation of hematopoietic progenitors, as well for prolonging survival of leukocytes, thus promoting both new cell production and extending survival of existing cells to achieve rapid increases in leukocyte numbers.

The pathways responsible for adult tissue homeostasis are governed significantly but not exclusively by Bcl-2–family proteins.2  The central pathway involved in daily programmed cell death in most tissues involves mitochondria, energy-producing organelles that play critical roles in both cell life and death.3  Several Bcl-2–family proteins, both antiapoptotic and proapoptotic, have C-terminal transmembrane domains that insert in the outer membrane of mitochondria. Proapoptotic Bcl-2–family proteins, such as Bax and Bak, induce mitochondrial outer membrane permeabilization (MOMP), causing the release of caspase-activating proteins and other cell death mediators, whereas antiapoptotic proteins such as Bcl-2 serve as guardians of the outer membrane and preserve its integrity by opposing Bax and Bak. Other nonmitochondrial pathways for apoptotic cell death also exist, including those governed by tumor necrosis factor-family death receptors, such as Fas—an important regulator of lymphoid homeostasis in vivo. However, even the death receptor pathway (“extrinsic pathway”) converges with the mitochondrial pathway (“intrinsic pathway”) in certain types of cells, through caspase-mediated cleavage and activation of Bid, an endogenous modulator of Bcl-2/Bax–family proteins.4 

Although mitochondria clearly induce apoptosis by releasing proteins that participate in caspase activation (eg, cytochrome c) and that neutralize endogenous inhibitors of caspases (eg, SMAC, OMI/Htra2), these organelles are also well known mediators of necrotic cell death.5  For example, defects in electron chain transport in respiring mitochondria spew reactive oxygen species into cells, causing lipid peroxidation and membrane damage, which impair normal ion-homeostasis, causing cellular swelling and plasma membrane rupture, as well as rupture of lysosomes and release of hydrolytic enzymes that destroy proteins, nucleic acids, and lipids. Bcl-2–family proteins also modulate these necrotic actions of mitochondria.6  The point of regulation may be linked to the ability of Bcl-2–family proteins to control outer mitochondrial membrane permeability, inasmuch as loss of cytochrome c from mitochondria interrupts electron chain transport between complexes III and IV (although this notion has been challenged).7  In addition, once downstream caspases are activated, they can cleave proteins required for proper function of complex I.8  MOMP also releases several proteins that contribute to nonapoptotic cell death, including DNAse, endonuclease G, and apoptosis-inducing factor, a flavoprotein reported to enter the nucleus and promote genome destruction.9  Alternatively, Bcl-2–family proteins may have other, as yet poorly understood ways of communicating with the mitochondrial inner membrane to affect mitochondrial bioenergetics and control nonapoptotic cell death.10  The mechanistic details notwithstanding, the bottom line is that antiapoptotic proteins such as Bcl-2 protect and proapoptotic proteins such as Bax kill even when caspases are neutralized using broad-spectrum chemical inhibitors, showing that Bcl-2–family proteins control a cell death checkpoint upstream of caspase activation, thus allowing them to govern both apoptotic (caspase-dependent) and nonapoptotic (caspase-independent) cell death.

Antiapoptotic Bcl-2–family proteins are well known for their ability to prolong survival of growth factor–dependent cells when deprived of their obligate growth factors. In fact, the antiapoptotic function of Bcl-2 was first elucidated in gene transfection studies, where interleukin-3–dependent murine hematopoietic cells were shown to cease division but to survive for prolonged periods in the absence of interleukin-3 when Bcl-2 was constitutively overexpressed.11  Growth factor deprivation can also lead to nutrient deprivation, because expression of cell-surface glucose transporters and amino-acid transporters depends on them. Prolonged nutrient deprivation invokes autophagy, an evolutionarily conserved response for catabolizing macromolecules and organelles, thereby generating substrates for ATP production.12,13  Autophagy is initially induced to prolong cell survival, but when taken to extremes, it seems to cause cell death. Bcl-2 and Bcl-XL suppress autophagy by binding the protein Beclin (ATG7),14,15  an essential component of the mammalian autophagy system that marks autophagic vesicles for fusion with lysosomes for digestion and recycling of components.

The antiautophagic function of Bcl-2 has been dissociated from mitochondrial location, and instead appears to be manifested from the endoplasmic reticulum (ER),14  where a considerable proportion of antiapoptotic Bcl-2 and related proteins often resides.16  In this regard, Bcl-2–family proteins have regulatory effects on several proteins and processes in the ER, including those influencing the unfolded protein response (UPR).17  The UPR is an evolutionarily conserved adaptive mechanism that detects accumulation of unfolded proteins in the lumen of the ER, causing induction of chaperones and transporters that strive to restore homeostasis, at least in part, by retrograde transport of unfolded proteins into the cytosol for ubiquitination and proteasome-mediated destruction.18  Given that chaperone-mediated autophagy complements proteasome-dependent destruction of misfolded proteins, representing another major route of disposal of defective proteins,19  it is tempting to speculate that Bcl-2–family proteins have evolved additional functions that serve during times of cell stress to tip the balance toward either survival or death. Indeed, persistent ER stress induces both caspase-dependent and caspase-independent cell death programs, which are modulated by Bcl-2–family proteins.17  Altogether, therefore, Bcl-2–family proteins govern cellular pathways involved in apoptosis, necrosis, and autophagy (Figure 1).

Figure 1

Bcl-2 suppresses apoptosis, necrosis, and autophagy. ROS indicates reactive oxygen species; Cyt-c, cytochrome-c; EndoG, endonuclease G; AIF, apoptosis-inducing factor; and IAP, inhibitor of apoptosis protein.

Figure 1

Bcl-2 suppresses apoptosis, necrosis, and autophagy. ROS indicates reactive oxygen species; Cyt-c, cytochrome-c; EndoG, endonuclease G; AIF, apoptosis-inducing factor; and IAP, inhibitor of apoptosis protein.

Close modal

As we commemorate the 50th anniversary of Blood, it seems fitting to take a historical approach to the topic of Bcl-2–family proteins and their roles in hematology. Moreover, a brief historical accounting of some of the milestones in Bcl-2 research serves to illustrate several of the important functions and mechanisms of these proteins. Many of the first-time observations concerning Bcl-2 and related proteins were made in the context of the hematopoietic and lymphoid systems or malignancies originating from these cells. Even with the hematology-centric focus of this review, however, space limitations preclude citing all contributions and all contributors, for which I apologize in advance.

BCL-2 gene activation in hematopoietic malignancies

Cloning of portions of the human BCL-2 gene was first reported in 1985 by Tsujimoto et al,20  who cloned the breakpoints from t(14;18) chromosomal translocations observed in non-Hodgkin lymphomas. In these translocations, the BCL-2 gene becomes fused with the immunoglobulin heavy-chain locus (IgH), bringing the juxtaposed BCL-2 gene under the control of the IgH enhancer and thereby dysregulating BCL-2 gene expression at a transcriptional level. The somatic hypermutation mechanism associated with the IgH gene locus often then peppers the BCL-2 gene with mutations, which may further dysregulate expression and can also lead to point mutations in the coding regions of the Bcl-2 protein,21  some of which have functional significance but are not required for suppressing apoptosis. In 1986, successful cDNA cloning of Bcl-2 transcripts was accomplished by Tsujimoto et al22  and shortly thereafter by Cleary et al,23  allowing the amino-acid sequence of the Bcl-2 protein to be deduced. At the time, the only recognized homology to other proteins was BHRF, an open reading frame in Epstein-Barr virus (EBV),23  an observation that years later would be validated by various studies showing that EBV and several Herpes-family viruses contain functional Bcl-2 homologs.24  In 1992, Hanada et al25  described high levels of BCL-2 gene expression in most chronic lymphocytic leukemias (CLLs), in association with gene hypomethylation. Not until 2005, however, would it be revealed by Cimmino et al26  that BCL-2 gene expression in B-cells is normally repressed by endogenous microRNAs (miRs), and the genes encoding these regulatory, noncoding RNAs (miR15, miR16) become deleted or inactivated by mutations in more than 70% of CLLs. This discovery of loss of miR-mediated repression of BCL-2 represented the first clear elucidation of a miR-dependent mechanism of tumor suppression in humans. In addition to chromosomal translocations as a mechanism for activation of the BCL-2 gene in human malignancies, in 1997, examples of BCL-2 gene amplification were found in non-Hodgkin lymphomas.27 

Cytoprotective function of Bcl-2

In 1988, the first demonstration of oncogenic potential of Bcl-2 was made by Reed et al,28  confirming that Bcl-2 is a bona fide protooncogene. Shortly thereafter, the antiapoptotic activity of Bcl-2 was discovered by Vaux et al,11  a milestone in research on Bcl-2 that laid a foundation for the next 2 decades of work. The first Bcl-2 transgenic mice were reported in 1989 by McDonnell et al,29  showing that constitutive overexpression of Bcl-2 in vivo causes polyclonal accumulation of B lymphocytes as a result of delayed programmed cell death, and recapitulating the histologic features of the human follicular lymphomas that commonly bear t(14;18) translocations. In 1990, the first demonstrations that knocking down Bcl-2 expression induces leukemia cell death and inhibits leukemic growth in vivo were made by Reed and colleagues, using antisense methods.30,31  In 1991, bcl-2 gene knockout mice were generated by Sentman et al,32  revealing a requirement for Bcl-2 for postnatal survival of lymphocytes and melanocytes, observations that would later influence selection of indications for testing of Bcl-2–targeted therapies in human clinical trials.

The concept that Bcl-2 controls an evolutionarily conserved pathway for regulating programmed cell death was introduced in 1992 by Vaux et al33  with their demonstration that ectopic expression of human Bcl-2 in mutant worms (Caenorhabditis elegans) could rescue a defect in programmed cell death. Later, in 1994, the defective gene responsible for the excessive cell death observed in these mutant worms (Ced9) was cloned by Hengartner and Horvitz,34  proving that at least some elements of the cellular machinery that controls apoptosis are highly conserved within the animal kingdom. This theme of evolutionary conservation of mechanisms was continued by Kane et al,35  who in 1993 reported that human Bcl-2 could even protect yeast (Saccharomyces cerevisiae) under some circumstances. This work, along with that of Zhong et al,36  also showed some of the early hints that Bcl-2 could suppress both apoptotic and necrotic cell death, focusing on neuronal cell death models.

Bcl-2 confers resistance to chemotherapy

BCL-2 gene rearrangements were associated with poor prognosis in large-cell non-Hodgkin lymphomas in 1989 by Yunis et al,37  an observation subsequently confirmed and extended to immunohistochemical detection of elevated Bcl-2 protein by several groups.38-41  Although implying that Bcl-2 altered the biologic behavior of lymphoma in adverse ways, the experimental link between Bcl-2 and resistance to cytotoxic anticancer drugs was directly made in reports by Miyashita and Reed in 199242  and 1993,43  showing that stable overexpression of Bcl-2 in various lymphoid cell lines conveyed resistance to a broad range of DNA-damaging drugs, antimicrotubule drugs, nucleoside analogs, and glucocorticoids and building on the work of Sentman et al,32  who showed that Bcl-2 could prevent cell death induced by multiple types of cell death stimuli in thymocytes. The findings implied that Bcl-2 operates at a distal point in a conserved cell death pathway used by most anticancer drugs, and revealed a novel form of chemoresistance, distinct from the previously identified mechanisms involving drug efflux, drug metabolism, drug inactivation, and related mechanisms. Shortly thereafter, Campos et al44,45  showed that high levels of Bcl-2 are associated with resistance of patients with acute myeloid leukemia (AML) to chemotherapy and that antisense oligodeoxynucleotides targeting Bcl-2 mRNA improve sensitivity of AML cells to arabinocytosine (AraC) in vitro. Together, these and many other clinical correlative studies strengthened the association between Bcl-2 and chemoresistance in hematolymphoid malignancies, with similar results seen for some solid tumors, such as prostate cancer.46,47  Then, in 1997, the first phase 1 clinical trial of a Bcl-2–targeted therapy was reported by Webb et al,48  a study of an antisense phosphorothioate oligodeoxy-nucleotide that showed promising activity in patients with refractory lymphomas and that raised hopes that neutralizing Bcl-2 could restore apoptosis sensitivity to malignant cells and promote their demise.

Mechanisms of action of Bcl-2–family proteins

Several members of the family contain a stretch of hydrophobic residues at their C terminus that has been variably reported to be important for their functions. In 1987 and 1989, reports by Tsujimoto et al49  and Chen-Levy et al50  groups, respectively, confirmed that Bcl-2 is an integral membrane protein, found in intracellular membranes. Then, in 1990, association of Bcl-2 with mitochondria was discovered by Hockenbery et al51  and the suppressive action of Bcl-2 on mitochondrial reactive oxygen species (ROS) production was elucidated in connection with programmed cell death. In 1994, the first in vitro reconstitution of a Bcl-2–dependent cell death pathway was achieved by Newmeyer et al52  using Xenopus laevis extracts to show that Bcl-2–mediated protection depends on mitochondria. Then, in 1995, Zamzami et al53  showed that the loss of mitochondrial membrane potential was an early event associated with apoptosis, which was suppressible by Bcl-2. Subsequently, in 1996, Liu et al54  and Li et al55  identified cytochrome c as the chief link between mitochondria and apoptosis, binding to and activating the cytoplasmic caspase-activating protein, Apaf1. Bcl-2 was first shown to suppress cytochrome c release from mitochondria in 1997 by Yang et al56  an observation quickly confirmed by many other groups. At about this time, Wolter et al57  and Nechushtan et al58  demonstrated dynamic trafficking of Bcl-2–family proteins between cytosol and mitochondria and elucidated a pathway whereby latent Bax in the cytosol undergoes conformational changes that induce its insertion into mitochondrial membranes, a process suppressed by Bcl-2.57,58 

In 1997, the 3-dimensional (3D) structure of Bcl-XL was determined by Muchmore et al59  and Fesik,60  revealing for the first time that several Bcl-2–family members share striking structural similarity to the pore-forming domains of bacterial toxins and sparking a series of subsequent studies that demonstrated an ability of Bcl-XL, Bcl-2, and Bax to form ion channels in synthetic membranes.61-63  It remains a mystery how Bcl-2 and Bax can have essentially the same 3D protein fold64  and yet have diametrically opposed effects on cell death. Probably the best correlate with their opposing activities is oligomerization of Bax in mitochondrial membranes (first demonstrated by Eskes et al65  and Wei et al66  in 2000), which coincides with MOMP. Bcl-2 does not oligomerize in mitochondrial membranes, but rather blocks Bax oligomerization. Wei et al66  showed that BH3-only protein Bid serves as an agonist of Bax oligomerization in 2000. In 2002, Kuwana et al67  reconstituted in synthetic liposomes the Bax pore, showing that Bax can be induced to oligomerize by Bid BH3 peptide, permeabilizing liposomes in vitro. Thus, Bax and structurally related proapoptotic proteins (Bak, Bok) are hypothesized to form proteinaceous or lipidic pores upon oligomerization in mitochondrial membranes, alone or in conjunction with resident mitochondrial proteins, thereby allowing escape of cytochrome c and other proteins from these organelles (reviewed by Reed68 ). In parallel, in 2001, gene ablation studies in mice by Wei et al69  demonstrated that either Bax or Bak is necessary for MOMP, establishing these pore-forming proapoptotic proteins as the distal elements controlling permeability of mitochondrial membranes and thus the governors of life/death decisions in the mammalian mitochondrial pathway for cell death (Figure 2).

Figure 2

Functional interactions among the types of mammalian Bcl-2–family proteins at mitochondrial membranes.

Figure 2

Functional interactions among the types of mammalian Bcl-2–family proteins at mitochondrial membranes.

Close modal

In parallel with the early observations concerning effects of Bcl-2–family proteins on mitochondrial membrane permeability, researchers studying Ced-9, the C elegans ortholog of Bcl-2, reported in 1997 that Ced-9 directly binds to and inhibits Ced-4, the caspase-activating Apaf-1 ortholog.70-72  However, this mechanism for inhibition of caspase-activating protein would not find its human analog until 2007, with the discovery by Bruey et al73  that Bcl-2 binds to and suppresses NLR-family member NALP1, an Apaf1-like activator of inflammatory caspases. In 1998, Liang et al74  described interaction of Bcl-2 with Beclin, a protein that later would be recognized as a critical component of the autophagy machinery of mammalian cells, thus linking Bcl-2 to suppression of autophagy.14 

Proapoptotic Bcl-2 family members as tumor suppressors

An important link between apoptosis promotion and tumor suppression was made by Miyashita et al75  and Miyashita and Reed76  in 1994 and 1995, respectively, with the discovery that p53 directly binds to and suppresses BAX gene expression, thereby revealing the first proapoptotic target of this important tumor suppressor. Since then, multiple proapoptotic members of the Bcl-2 family, including Bid, Puma, and Noxa, have been shown to be direct targets of p53, with the relative importance of any particular family member dependent on cell lineage and cellular context.77-79  The first examples of loss of function mutations in any proapoptotic Bcl-2–family gene were uncovered by studies of the human BAX gene in colon cancers by Rampino et al80  in 1997 and later observed also for hematopoietic malignancies.81  At about this time, studies of bax−/− mice by Yin et al82  revealed that this proapoptotic gene suppresses tumorigenesis in vivo, firmly establishing Bax as a tumor suppressor. More recently, cytosolic interactions of p53 with pro- and antiapoptotic Bcl-2–family proteins have been observed by Chipuk et al83  and Deng et al,84  respectively, directly modulating the activities of Bax and Bcl-2, and suggesting that p53 regulates the Bcl-2 family at both transcriptional and posttranscriptional levels.

As membership in the Bcl-2 family grew with discovery of additional pro- and antiapoptotic members in the mid-1990s, regions of conserved sequence homology were pinpointed, and given various rubrics, coalescing on the nomenclature proposed by Yin et al85  of Bcl-2 Homology (BH) domains. Today, we recognize up to 4 BH domains. All antiapoptotic members of the human Bcl-2 family contain BH1, BH2, BH3, and BH4. Mutagenesis studies of BH1 and BH2 in antiapoptotic Bcl-2 suggested importance for homo- and heterodimerization with Bax.85  The fundamental motif of proapoptotic proteins required for dimerization with antiapoptotic Bcl-2–family proteins was first identified in 1995 by Chittenden et al,86  later termed BH3 (Bcl-2 Homology domain 3). Subsequently, several proapoptotic proteins were discovered that share sequence homology only in the BH3 domain (“BH3 only” proteins) and that depend upon the BH3 domain for binding Bcl-2 or Bcl-XL, including Bid, Bim, Hrk, Bik, Puma, Noxa, Bcl-GS, Bmf, Bnf, Mule, and possibly Spike (reviewed by Strasser87 ). These BH3-only proteins uniformly operate as antagonists of antiapoptotic Bcl-2–family proteins in a BH3-dependent manner that correlates with BH3-dependent binding. The BH3-only proteins stand apart from a smaller group of proapoptotic proteins that contain BH1, BH2, and BH3 (sometimes termed “multidomain”), which includes Bax, Bak, and Bok in mammals. A few BH3-only proteins (ie, Bid, Bim, Puma) have dual functions as both antagonists of antiapoptotic Bcl-2 family members and agonists of proapoptotic multidomain proteins Bax and Bak.

Attempts to understand dimerization among Bcl-2 family proteins remained confusing until Muchmore et al59  solved the first 3D structure of a Bcl-2–family member, Bcl-XL, in 1996, and then elucidated the structure of a complex of Bcl-XL with the BH3 peptide from Bak in 1997.88  These seminal studies taught that the BH3 peptide is an amphipathic α-helix of approximately 16 to 20 amino acids that binds a hydrophobic pocket on the surface of antiapoptotic Bcl-2–family proteins, thus revealing the structural basis for antagonism of pro- and antiapoptotic members of the family and setting the stage for subsequent drug discovery strategies based on mimicking BH3 peptides with chemical compounds that bind the same pocket.89  The first chemical inhibitor of Bcl-2 based on this concept was described by Wang et al90  in 2000, who used computational modeling approaches based on the structural data. Since then, a variety of approaches have resulted in successful identification and synthesis of chemical antagonists of antiapoptotic Bcl-2 family proteins, some of which are now in clinical testing (reviewed by Reed and Pellecchia91 ).

Posttranslational modifications

Many proteins are regulated by posttranslational modifications, and Bcl-2 family proteins are no exception. To date, phosphorylation, proteolysis, ubiquitination, myristoylation, and deamidation are among the documented posttranslational modifications of Bcl-2–family proteins. The first report of phosphorylation of Bcl-2 was made in 1994 by May et al,92  and it was originally reported to be inhibitory, although stimulatory effects have since been described93 —an issue that remains unresolved. Phosphorylation of proapoptotic Bcl-2-family proteins can either neutralize (eg, BAD) or potentiate (eg, BIM) their functions. Far more work is needed on biochemical and functional analysis of posttranslational modifications of Bcl-2–family members.

Dual phenotype of Bcl-2

Bcl-2 is best known for its ability to suppress apoptosis, but hints that this protein could also participate in apoptosis induction began to surface as early as 1996.94  In 1997, Cheng et al95  revealed the first mechanism for conversion of Bcl-2 from a protector to a killer, showing that proteolytic removal of N-terminal sequences by caspase-mediated cleavage flips the phenotype of Bcl-2. It is noteworthy that mutations in translocated BCL-2 alleles have been identified in human lymphomas that ablate the aspartic acid residue required for caspase cleavage,21  suggesting that some tumors may evolve strategies for avoiding the “dark side” of Bcl-2. Later, in 2004, Lin et al96  discovered another mechanism for converting Bcl-2 into a killer, showing that the orphan nuclear receptor Nur77 can be induced to translocate from nucleus to cytosol, binding Bcl-2, and inducing a conformational change in Bcl-2 that probably mimics what happens during caspase cleavage, exposing the normally buried BH3 domain of Bcl-2 and causing it to function as a proapoptotic protein. A potentially similar mechanism was identified for Bcl-XL in 2006 by Bivona et al,97  showing that lipid modifications of K-Ras can promote its association with Bcl-XL on mitochondria and induce apoptosis. Thus, depending on the proteins with which Bcl-2 and Bcl-XL interact, their phenotypes can be converted from anti- to proapoptotic, revealing an additional level of complexity to these proteins that has clear therapeutic implications for malignancies that overexpress these Bcl-2-family proteins. Perhaps phenotypic conversion underlies the recent discovery of compounds by Schwartz et al98  that show superior cytotoxic activity in Bcl-2/Bcl-XL–overexpressing cells compared with control cells, as well as providing a potential explanation for why high levels of Bcl-2 expression are sometimes associated with better patient prognosis—first observed in breast cancers in 1994 by Silvestrini et al.99 

Unanswered questions

During the 2 decades since the discovery of Bcl-2, we have learned much about the functions of Bcl-2–family proteins. The lessons learned, however, have also taught us that there is much we do not understand, including (1) the structural details of how Bax and Bak permeabilize mitochondrial membranes, (2) the role of Bcl-2–family proteins in ER membranes, where they modulate Ca2+ and influence signal transduction events linked to cell cycle entry and ER stress,17,100  (3) the relevance of Bcl-2–mediated suppression of autophagy on tissue homeostasis and neoplasia, and (4) the many ways that posttranslational modifications (phosphorylation, deamidation, proteolysis) can alter the phenotypes of Bcl-2 family proteins.101 

In vivo actions of Bcl-2–family proteins—of mice and men

Although genetic alterations that activate the BCL-2 proto-oncogene or inactivate the BAX tumor suppressor gene in human malignancies have provided important clues about the in vivo roles of Bcl-2–family proteins in the hematopoietic system, the preponderance of what we know has come from investigations of genetically engineered mice. Space limitations allow only a superficial treatment of what we have learned about regulation of programmed cell death in the hematopoietic system from mouse genetics studies, and thus the reader is referred to recent reviews on the topic by Strasser87  and Opferman and Korsmeyer.102  In brief, in terms of basic tissue homeostasis, gene knockout studies have shown that (1) Bcl-2 is required for maintenance of peripheral lymphocyte populations in adulthood, with animals developing lymphopenias at approximately the time they undergo sexual maturation (probably a time when endogenous glucocorticoid levels rise)103 ; (2) Bax-deficient mice develop lymphocytosis, demonstrating a role for this proapoptotic protein in programmed cell death of mature lymphocytes104 ; (3) Bcl-X is required for survival of immature thymocytes105 ; (4) Bid knockout mice are initially hematologically normal but after a long latency period develop chronic myelomonocytic leukemias, suggesting a role for Bid in controlling turnover of some populations of myeloid cells106 ; (5) proapoptotic gene Bim is required for proper control of granulocyte survival, elimination of autoreactive thymocytes, and programmed cell death of memory B cells and plasma cells (or their progenitors)107-110 ; (6) antiapoptotic Mcl-1 is required for survival of hematopoietic stem cells, survival of developing T and B lymphocytes, as well as maintenance of mature lymphocyte cell populations111,112 ; (7) proapoptotic Bak helps to govern the homeostasis of B-cell populations in vivo and bak−/− platelets have extended half-life113 ; (8) antiapoptotic Bcl-W is expressed in myeloid cells but evidently not essential for survival114 ; (9) Bad-deficient mice develop diffuse large B-cell lymphoma, suggesting a tumor suppressor role for this proapoptotic protein in mature B cells115 ; and (10) neutrophils and mast cells from mice deficient in A1a (murine ortholog of Bfl-1) show accelerated apoptosis in culture.116,117  We have also learned from gene knockout studies in mice that defects in programmed cell death promote autoimmunity, presumably by interfering with eradication of autoreactive lymphocytes or their progenitors.

Species-specific differences in some members of the Bcl-2 gene families of humans and mice have made it difficult to extend the in vivo analysis to all members. For example, in humans, the nuclear factor-κB–inducible antiapoptotic gene BFL-1 is represented by a single gene in humans, but consists of 4 highly homologous genes spread around the mouse genome (A1a, A1b, A1c, A1d), making it difficult to perform genetic ablation experiments.118  In addition, in mice, the closest homolog to the human BCL-B (BCL210L) gene is diva/boo, which has entirely different patterns of expression in vivo in humans versus mice, with expression restricted to reproductive tissues in mice but prominent in plasma cell and some populations of B lymphocytes (among other types of cells) in humans.119  Thus, although the hematopoietic systems of diva/boo knockout mice are phenotypically normal,120  we cannot infer that the same would be true in humans lacking the corresponding gene.

Bcl-2–family proteins as central integrators—critical nodes in life's networks

Although much research has focused on homo- and heterodimerization among Bcl-2–family proteins, many if not most of these proteins have other interaction partners that regulate their activity and that link them to a wide variety of cellular pathways, giving the impression that Bcl-2–family proteins operate as critical nodes in complex networks to integrate information and make ultimate life/death decisions. For the more heavily studied members of the family, such as Bcl-2, Bcl-XL, and Bax, the list of potential interacting proteins is almost overwhelming (Figure 3). The structural basis for most of these nonfamilial protein interactions has not been elucidated, and their roles in physiology are not always well understood, but the extensive repertoire of partners suggests that diverse signaling, developmental, and metabolic pathways converge on Bcl-2–family members.

Figure 3

Nonfamilial interactions with antiapoptotic proteins Bcl-2 or Bcl-XL and with proapoptotic proteins Bax or Bak. (A) The depicted protein interactions have been reported for either Bcl-2 or Bcl-XL. (B) The depicted protein interactions have been reported for Bax or Bak. In many cells, Bax is present in a latent (inactive) conformation in the cytosol,57  in which its C-terminal transmembrane domain is folded onto the protein.64  Activators of Bax induce conformational changes that promote Bax's insertion into membranes of mitochondria, followed by BH3-induced oligomerization in membranes. Several proteins that modulate these steps have been reported.

Figure 3

Nonfamilial interactions with antiapoptotic proteins Bcl-2 or Bcl-XL and with proapoptotic proteins Bax or Bak. (A) The depicted protein interactions have been reported for either Bcl-2 or Bcl-XL. (B) The depicted protein interactions have been reported for Bax or Bak. In many cells, Bax is present in a latent (inactive) conformation in the cytosol,57  in which its C-terminal transmembrane domain is folded onto the protein.64  Activators of Bax induce conformational changes that promote Bax's insertion into membranes of mitochondria, followed by BH3-induced oligomerization in membranes. Several proteins that modulate these steps have been reported.

Close modal

Future therapeutic opportunities

The fruits of more than 2 decades of research on Bcl-2 and its kin have yielded new strategies for therapeutic applications, some of which have advanced to clinical testing in humans. The first of these is antisense oligonucleotides targeting Bcl-2 mRNA, which have shown promising activity for CLL in randomized phase 3 trials.121  Results for myeloma and AML have not been encouraging, which could be due to the overexpression of antiapoptotic members of the family besides Bcl-2. Multiple chemical antagonists of antiapoptotic Bcl-2–family proteins have been described that bind the same pocket occupied by proapoptotic BH3 domains (reviewed by Reed and Pellecchia91 ). These compounds have different potencies and variable spectra of activity against the 6 antiapoptotic members. At least 3 such compounds have advanced into human testing to date, with more likely to follow soon. Compounds that appear to convert Bcl-2, Bcl-XL, or their relatives from protectors to killers have been identified, including molecules that directly bind these proteins and others that trigger Nur77 translocation.122  A close analog of a Nur77-modulating compound is in clinical testing currently for cancer treatment. These translational aspects of research on Bcl-2 raise hopes that a new class of anticancer drugs may be forthcoming. Given the prominent role of Bcl-2–family proteins in elimination of autoreactive lymphocytes, these experimental therapeutic agents are likely to also find applications for a variety of autoimmune diseases.

I thank M. Hanaii and T. Siegfried for manuscript preparation.

This work was supported by the National Institutes of Health NCI-Drug Discovery Group (CA113318) and a grant on apoptosis and mitochondria (GM60554).

National Institutes of Health

Contribution: J.C.R. is the sole author.

Conflict-of-interest disclosure: The author declares no competing financial interests.

Correspondence: John C. Reed, 10901 N Torrey Pines Road, Burnham Institute for Medical Research, La Jolla, CA 92037; e-mail: reedoffice@burnham.org.

1
Salvesen
 
GS
Abrams
 
JM
Caspase activation—stepping on the gas or releasing the brakes? Lessons from humans and flies.
Oncogene
2004
, vol. 
23
 (pg. 
2774
-
2784
)
2
Cory
 
S
Huang
 
DC
Adams
 
JM
The Bcl-2 family: roles in cell survival and oncogenesis.
Oncogene
2003
, vol. 
22
 (pg. 
8590
-
8607
)
3
Green
 
DR
Kroemer
 
G
The pathophysiology of mitochondrial cell death.
Science
2004
, vol. 
305
 (pg. 
626
-
629
)
4
Korsmeyer
 
SJ
Wei
 
MC
Saito
 
M
Weiler
 
S
Oh
 
KJ
Schlesinger
 
PH
Proapoptotic cascade activates BID, which oligomerizes BAK or BAX into pores that result in the release of cytochrome c.
Cell Death Differ
2000
, vol. 
7
 (pg. 
1166
-
1173
)
5
Reed
 
JC
Apoptosis-based therapies.
Nat Rev Drug Discov
2002
, vol. 
1
 (pg. 
111
-
121
)
6
Bredesen
 
DE
Rao
 
RV
Mehlen
 
P
Cell death in the nervous system.
Nature
2006
, vol. 
443
 (pg. 
796
-
802
)
7
Waterhouse
 
NJ
Goldstein
 
JC
von Ahsen
 
O
Schuler
 
M
Newmeyer
 
DD
Green
 
DR
Cytochrome c maintains mitochondrial transmembrane potential and ATP generation after outer mitochondrial membrane permeabilization during the apoptotic process.
J Cell Biol
2001
, vol. 
153
 (pg. 
319
-
328
)
8
Ricci
 
JE
Munoz-Pinedo
 
C
Fitzgerald
 
P
, et al. 
Disruption of mitochondrial function during apoptosis is mediated by caspase cleavage of the p75 subunit (NDUSF1) of Complex I of the electron transport chain.
Cell
2004
, vol. 
117
 (pg. 
773
-
86
)
9
Penninger
 
JM
Kroemer
 
G
Mitochondria, AIF and caspases–rivaling for cell death execution.
Nat Cell Biol
2003
, vol. 
5
 (pg. 
97
-
99
)
10
Reed
 
JC
Green
 
DR
Remodeling for demolition: Changes in mitochondrial ultrastructure during apoptosis.
Mol Cell
2002
, vol. 
9
 (pg. 
1
-
3
)
11
Vaux
 
DL
Cory
 
S
Adams
 
JM
Bcl-2 gene promotes haemopoietic cell survival and cooperates with c-myc to immortalize pre-B cells.
Nature
1988
, vol. 
335
 (pg. 
440
-
442
)
12
Edinger
 
AL
Thompson
 
CB
Death by design: apoptosis, necrosis and autophagy.
Curr Opin Cell Biol
2004
, vol. 
16
 (pg. 
663
-
669
)
13
Lum
 
JJ
DeBerardinis
 
RJ
Thompson
 
CB
Autophagy in metazoans: cell survival in the land of plenty.
Nat Rev Mol Cell Biol
2005
, vol. 
6
 (pg. 
439
-
448
)
14
Pattingre
 
S
Tassa
 
A
Qu
 
X
, et al. 
Bcl-2 antiapoptotic proteins inhibit beclin 1- dependent autophagy.
Cell
2005
, vol. 
122
 (pg. 
927
-
939
)
15
Shimizu
 
S
Kanaseki
 
T
Mizushima
 
N
, et al. 
Role of Bcl-2 family proteins in a non-apoptotic programmed cell death dependent on autophagy genes.
Nat Cell Biol
2004
, vol. 
6
 (pg. 
1221
-
1228
)
16
Krajewski
 
S
Tanaka
 
S
Takayama
 
S
Schibler
 
MJ
Fenton
 
W
Reed
 
JC
Investigation of the subcellular distribution of the bcl-2 oncoprotein: residence in the nuclear envelope, endoplasmic reticulum, and outer mitochondrial membranes.
Cancer Res
1993
, vol. 
53
 (pg. 
4701
-
4714
)
17
Xu
 
C
Bailly-Maitre
 
B
Reed
 
JC
Endoplamic reticulum stress: Cell life and death decisions.
J Clin Invest
2005
, vol. 
115
 (pg. 
2656
-
2664
)
18
Ron
 
D
Cell biology. Stressed cells cope with protein overload.
Science
2006
, vol. 
313
 (pg. 
52
-
53
)
19
Kitsis
 
RN
Peng
 
C-F
Cuervo
 
AM
Eat your heart out.
Nat Med
2007
, vol. 
13
 (pg. 
539
-
541
)
20
Tsujimoto
 
Y
Cossman
 
J
Jaffe
 
E
Croce
 
C
Involvement of the Bcl-2 gene in human follicular lymphoma.
Science
1985
, vol. 
228
 (pg. 
1440
-
1443
)
21
Tanaka
 
S
Louie
 
D
Kant
 
J
Reed
 
J
Frequent somatic mutations in translocated BcL2 genes of non-Hodgkin's lymphoma patients.
Blood
1992
, vol. 
79
 (pg. 
229
-
237
)
22
Tsujimoto
 
Y
Croce
 
C
Analysis of the structure, transcripts, and protein products of Bcl-2, the gene involved in human follicular lymphoma.
Proc Natl Acad Sci U S A
1986
, vol. 
83
 (pg. 
5214
-
5218
)
23
Cleary
 
ML
Smith
 
SD
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
1986
, vol. 
47
 (pg. 
19
-
28
)
24
Henderson
 
S
Huen
 
D
Rowe
 
M
Dawson
 
C
Johnson
 
G
Rickinson
 
A
Epstein-Barr virus-coded BHRF1 protein, a viral homologue of Bcl-2, protects human B cells from programmed cell death.
Proc Natl Acad Sci U S A
1993
, vol. 
90
 (pg. 
8479
-
8483
)
25
Hanada
 
M
Delia
 
D
Aiello
 
A
Stadtmauer
 
E
Reed
 
J
Bcl-2 gene hypomethylation and high-level expression in B-cell chronic lymphocytic leukemia.
Blood
1993
, vol. 
82
 (pg. 
1820
-
1828
)
26
Cimmino
 
A
Calin
 
GA
Fabbri
 
M
, et al. 
miR-15 and miR-16 induce apoptosis by targeting BCL2.
Proc Nat Acad Sci U S A
2005
, vol. 
102
 (pg. 
13944
-
13949
)
27
Monni
 
O
Joensuu
 
H
Franssila
 
K
Klefstrom
 
J
Alitalo
 
K
Knuutila
 
S
BCL2 overexpression associated with chromosomal amplification in diffuse large B-cell lymphoma.
Blood
1997
, vol. 
90
 (pg. 
1168
-
1174
)
28
Reed
 
JC
Cuddy
 
M
Slabiak
 
T
Croce
 
CM
Nowell
 
PC
Oncogenic potential of bcl-2 demonstrated by gene transfer.
Nature
1988
, vol. 
336
 (pg. 
259
-
261
)
29
McDonnell
 
TJ
Deane
 
N
Platt
 
FM
, et al. 
Bcl-2-immunoglobulin transgenic mice demonstrate extended B-cell survival and follicular lymphoproliferation.
Cell
1989
, vol. 
57
 (pg. 
79
-
88
)
30
Reed
 
JC
Stein
 
C
Subasinghe
 
C
, et al. 
Antisense-mediated inhibition of BCL2 proto-oncogene expression and leukemic cell growth and survival: comparisons of phosphodiester and phosphorothioate oligodeoxynucleotides.
Cancer Res
1990
, vol. 
50
 (pg. 
6565
-
6570
)
31
Reed
 
JC
Cuddy
 
M
Haldar
 
S
, et al. 
BCL2-mediated tumorigenicity of a human Tlymphoid cell line: synergy with MYC and inhibition by BCL2 antisense.
Proc Natl Acad Sci USA
1990
, vol. 
87
 (pg. 
3660
-
3664
)
32
Sentman
 
CL
Shutter
 
JR
Hockenbery
 
D
Kanagawa
 
O
Korsmeyer
 
SJ
Bcl-2 inhibits multiple forms of apoptosis but not negative selection in thymocytes.
Cell
1991
, vol. 
67
 (pg. 
879
-
888
)
33
Vaux
 
D
Weissman
 
I
Kim
 
S
Prevention of programmed cell death in Caenorhabditis elegans by human Bcl-2.
Science
1992
, vol. 
258
 (pg. 
1955
-
1957
)
34
Hengartner
 
MO
Horvitz
 
HR
C. elegans cell survival gene ced-9 encodes a functional homolog of the mammlian proto-oncogene Bcl-2.
Cell
1994
, vol. 
76
 (pg. 
665
-
676
)
35
Kane
 
DJ
Sarafin
 
TA
Auton
 
S
, et al. 
Bcl-2 inhibition of neural cell death: decreased generation of reactive oxygen species.
Science
1993
, vol. 
262
 (pg. 
1274
-
1276
)
36
Zhong
 
LT
Sarafian
 
T
Kane
 
DJ
, et al. 
Bcl-2 inhibits death of central neural cells induced by multiple agents.
Proc Natl Acad Sci U S A
1993
, vol. 
90
 (pg. 
4533
-
4537
)
37
Yunis
 
JJ
Mayer
 
MG
Arensen
 
MA
Aeppli
 
DP
Oken
 
MM
Frizzera
 
G
Bcl-2 and other genomic alterations in the prognosis of large-cell lymphomas.
N Engl J Med
1989
, vol. 
320
 (pg. 
1047
-
1054
)
38
Hermine
 
O
Haioun
 
C
Lepage
 
E
, et al. 
Prognostic significance of Bcl-2 protein expression in aggressive non-Hodgkin's lymphoma.
Blood
1996
, vol. 
87
 (pg. 
265
-
272
)
39
Tang
 
SC
Visser
 
L
Hepperle
 
B
Hanson
 
J
Poppema
 
S
Clinical significance of bcl-2-MBR gene rearrangement and protein expression in diffuse large-cell non-Hodgkin's lymphoma: an analysis of 83 cases.
J Clin Oncol
1994
, vol. 
12
 (pg. 
149
-
154
)
40
Hill
 
ME
MacLennan
 
KA
Cunningham
 
DC
, et al. 
Prognostic significance of BCL-2 expression and bcl-2 major breakpoint region rearrangement in diffuse large cell non-Hodgkin's lymphoma: a British national lymphoma investigation study.
Blood
1996
, vol. 
88
 (pg. 
1046
-
1051
)
41
Gascoyne
 
RD
Adomat
 
SA
Krajewski
 
S
, et al. 
Prognostic significance of Bcl-2 protein expression and Bcl-2 gene rearrangement in diffuse aggressive non-Hodgkin's lymphoma.
Blood
1997
, vol. 
90
 (pg. 
244
-
251
)
42
Miyashita
 
T
Reed
 
JC
Bcl-2 oncoprotein blocks chemotherapy-induced apoptosis in a human leukemia cell line.
Blood
1993
, vol. 
81
 (pg. 
151
-
157
)
43
Miyashita
 
T
Reed
 
JC
Bcl-2 gene transfer increases relative resistance of S49.1 and WEHI7.2 lymphoid cells to cell death and DNA fragmentation induced by glucocorticoids and multiple chemotherapeutic drugs.
Cancer Res
1992
, vol. 
52
 (pg. 
5407
-
5411
)
44
Campos
 
L
Roualult
 
J-P
Sabido
 
O
, et al. 
High expression of Bcl-2 protein in acute myeloid leukemia cells is associated with poor response to chemotherapy.
Blood
1993
, vol. 
81
 (pg. 
3091
-
3096
)
45
Campos
 
L
Sabido
 
O
Rouault
 
J-P
Guyotat
 
D
Effects of Bcl-2 antisense oligodeoxynucleotides on in vitro proliferation and survival of normal marrow progenitors and leukemic cells.
Blood
1994
, vol. 
84
 (pg. 
595
-
600
)
46
Colombel
 
M
Symmans
 
F
Gil
 
S
, et al. 
Detection of the apoptosis-suppressing oncoprotein Bcl-2 in hormone-refractory human prostate cancers.
Am J Pathol
1993
, vol. 
143
 (pg. 
390
-
400
)
47
McDonnell
 
TJ
Troncoso
 
P
Brisbay
 
SM
, et al. 
Expression of the proto-oncogene Bcl-2 in the prostate and its association with emergence of androgen-independent prostate cancer.
Cancer Res
1992
, vol. 
52
 (pg. 
6940
-
6944
)
48
Webb
 
A
Cunningham
 
D
Cotter
 
F
, et al. 
BCL-2 antisense therapy in patients with non-Hodgkin lymphoma.
Lancet
1997
, vol. 
349
 (pg. 
1137
-
1141
)
49
Tsujimoto
 
Y
Ikegaki
 
N
Croce
 
CM
Characterization of the protein product of Bcl-2, the gene involved in human follicular lymphoma.
Oncogene
1987
, vol. 
2
 (pg. 
3
-
7
)
50
Chen-Levy
 
S
Nourse
 
J
Cleary
 
MZ
The Bcl-2 candidate proto-oncogene product is a 24-KD integral membrane protein highly expressed in lymphoid cell lines and lymphomas carrying the t(14;18) translocation.
Mol Cell Biol
1989
, vol. 
9
 (pg. 
701
-
710
)
51
Hockenbery
 
DM
Nunez
 
G
Milliman
 
C
Schreiber
 
RD
Korsmeyer
 
SJ
Bcl-2 is an inner mitochondrial membrane protein that blocks programmed cell death.
Nature
1990
, vol. 
348
 (pg. 
334
-
336
)
52
Newmeyer
 
D
Farschon
 
DM
Reed
 
JC
Cell-free apoptosis in Xenopus egg extracts: inhibition by Bcl-2 and requirement for an organelle fraction enriched in mitochondria.
Cell
1994
, vol. 
79
 (pg. 
353
-
364
)
53
Zamzami
 
N
Marchetti
 
P
Castedo
 
M
, et al. 
Reduction in mitochondrial potential constitutes an early irreversible step of programmed lymphocyte death in vivo.
J Exp Med
1995
, vol. 
181
 (pg. 
1661
-
1672
)
54
Liu
 
X
Kim
 
CN
Yang
 
J
Jemmerson
 
R
Wang
 
X
Induction of apoptotic program in cell-free extracts: requirement for dATP and cytochrome c.
Cell
1996
, vol. 
86
 (pg. 
147
-
157
)
55
Li
 
P
Nijhawan
 
D
Budihardjo
 
I
, et al. 
Cytochrome c and dATP-dependent formation of Apaf-1/Caspase-9 complex initiates an apoptotic protease cascade.
Cell
1997
, vol. 
91
 (pg. 
479
-
489
)
56
Yang
 
J
Liu
 
X
Bhalla
 
K
, et al. 
Prevention of apoptosis by Bcl-2: release of cytochrome c from mitochondria blocked.
Science
1997
, vol. 
275
 (pg. 
1129
-
1132
)
57
Wolter
 
KG
Hsu
 
YT
Smith
 
CL
Nechushtan
 
A
Xi
 
XG
Youle
 
RJ
Movement of bax from the cytosol to mitochondria during apoptosis.
J Cell Biol
1997
, vol. 
139
 (pg. 
1281
-
1292
)
58
Nechushtan
 
A
Smith
 
C
Hsu
 
YT
Youle
 
R
Conformation of the Bax C-terminus regulates subcellular location and cell death.
EMBO J
1999
, vol. 
18
 (pg. 
2330
-
2341
)
59
Muchmore
 
SW
Sattler
 
M
Liang
 
H
, et al. 
X-ray and NMR structure of human Bcl-XL, an inhibitor of programmed cell death.
Nature
1996
, vol. 
381
 (pg. 
335
-
341
)
60
Fesik
 
SW
Insights into programmed cell death through structural biology.
Cell
2000
, vol. 
103
 (pg. 
273
-
282
)
61
Minn
 
AJ
Velez
 
P
Schendel
 
SL
, et al. 
Bcl-xL forms an ion channel in synthetic lipid membranes.
Nature
1997
, vol. 
385
 (pg. 
353
-
357
)
62
Schendel
 
SL
Xie
 
Z
Montal
 
MO
Matsuyama
 
S
Montal
 
M
Reed
 
JC
Channel formation by antiapoptotic protein Bcl-2.
Proc Natl Acad Sci U S A
1997
, vol. 
94
 (pg. 
5113
-
5118
)
63
Antonsson
 
B
Conti
 
F
Ciavatta
 
A
, et al. 
Inhibition of Bax channel-forming activity by Bcl-2.
Science
1997
, vol. 
277
 (pg. 
370
-
372
)
64
Suzuki
 
M
Youle
 
RJ
Tjandra
 
N
Structure of bax: Coregulation of dimmer formation and intracellular localization.
Cell
2000
, vol. 
103
 (pg. 
645
-
654
)
65
Eskes
 
R
Desagher
 
S
Antonsson
 
B
Martinou
 
JC
Bid induces the oligomerization and insertion of Bax into the outer mitochondrial membrane.
Mol Cell Biol
2000
, vol. 
20
 (pg. 
929
-
935
)
66
Wei
 
MC
Lindsten
 
T
Mootha
 
VK
, et al. 
tBID, a membrane-targeted death ligand, oligomerizes BAK to release cytochrome c.
Genes Dev
2000
, vol. 
14
 (pg. 
2060
-
2071
)
67
Kuwana
 
T
Mackey
 
MR
Perkins
 
G
, et al. 
Bid, bax, and lipids cooperate to form supramolecular openings in the outer mitochondrial membrane.
Cell
2002
, vol. 
111
 (pg. 
331
-
342
)
68
Reed
 
JC
Proapoptotic multidomain Bcl-2/Bax-family proteins: mechanisms, physiological roles, and therapeutic opportunities.
Cell Death Differ
2006
, vol. 
13
 (pg. 
1378
-
1386
)
69
Wei
 
MC
Zong
 
WX
Cheng
 
EH
, et al. 
Proapoptotic BAX and BAK: a requisite gateway to mitochondrial dysfunction and death.
Science
2001
, vol. 
292
 (pg. 
727
-
730
)
70
Chinnaiyan
 
AM
O'Rourke
 
K
Lane
 
BR
Dixit
 
VM
Interaction of CED-4 with CED-3 and CED-9: a molecular framework for cell death.
Science
1997
, vol. 
275
 (pg. 
1122
-
1126
)
71
Spector
 
MS
Desnoyers
 
S
Heoppner
 
DJ
Hengartner
 
MO
Interaction between the C. elegans cell-death regulators CED-9 and CED-4.
Nature
1997
, vol. 
385
 (pg. 
653
-
656
)
72
Wu
 
D
Wallen
 
H
Inohara
 
N
Nunez
 
G
Interaction and regulation of the c. elegans death protease CED-3 by CED-4 and CED-9.
J Biol Chem
1997
, vol. 
272
 (pg. 
21449
-
21454
)
73
Bruey
 
JM
Bruey-Sedano
 
N
Luciano
 
F
, et al. 
Bcl-2 and Bcl-XL regulate proinflammatory caspase-1 activation by interaction with NALP1.
Cell
2007
, vol. 
129
 (pg. 
45
-
56
)
74
Liang
 
XH
Kleeman
 
LK
Jiang
 
HH
, et al. 
Protection against fatal Sindbis virus encephalitis by beclin, a novel Bcl-2-interacting protein.
J Virol
1998
, vol. 
72
 (pg. 
8586
-
8596
)
75
Miyashita
 
T
Krajewski
 
S
Krajewska
 
M
, et al. 
Tumor suppressor p53 is a regulator of BCL-2 and BAX in gene expression in vitro and in vivo.
Oncogene
1994
, vol. 
9
 (pg. 
1799
-
1805
)
76
Miyashita
 
T
Reed
 
JC
Tumor suppressor p53 is a direct transcriptional activator of human Bax gene.
Cell
1995
, vol. 
80
 (pg. 
293
-
299
)
77
Oda
 
E
Ohki
 
R
Murasawa
 
H
, et al. 
Noxa, a BH3-only member of the Bcl-2 family and candidate mediator of p53-induced apoptosis.
Science
2000
, vol. 
288
 (pg. 
1053
-
1058
)
78
Sax
 
JK
Fei
 
P
Murphy
 
ME
Bernhard
 
E
Korsmeyer
 
SJ
El-Deiry
 
WS
BID regulation by p53 contributes to chemosensitivity.
Nat Cell Biol
2002
, vol. 
4
 (pg. 
842
-
849
)
79
Yu
 
J
Wang
 
Z
Kinzler
 
KW
Vogelstein
 
B
Zhang
 
L
PUMA mediates the apoptotic response to p53 in colorectal cancer cells.
Proc Natl Acad Sci U S A
2003
, vol. 
100
 (pg. 
1931
-
1936
)
80
Rampino
 
N
Yamamoto
 
H
Ionov
 
Y
, et al. 
Somatic frameshift mutations in the BAX gene in colon cancers of the microsatellite mutator phenotype.
Science
1997
, vol. 
275
 (pg. 
967
-
969
)
81
Meijerink
 
JP
Mensink
 
EJ
Wang
 
K
, et al. 
Hematopoietic malignancies demonstrate loss-of-function mutations of BAX.
Blood
1998
, vol. 
91
 (pg. 
2991
-
2997
)
82
Yin
 
C
Knudson
 
CM
Korsmeyer
 
SJ
Van Dyke
 
T
Bax suppresses tumorigenesis and stimulates apoptosis in vivo.
Nature
1997
, vol. 
385
 (pg. 
637
-
640
)
83
Chipuk
 
JE
Kuwana
 
T
Bouchier-Hayes
 
L
, et al. 
Direct activation of Bax by p53 mediates mitochondrial membrane permeabilization and apoptosis.
Science
2004
, vol. 
303
 (pg. 
1010
-
1014
)
84
Deng
 
X
Gao
 
F
Flagg
 
T
Anderson
 
J
May
 
WS
Bcl2's flexible loop domain regulates p53 binding and survival.
Mol Cell Biol
2006
, vol. 
26
 (pg. 
4421
-
4434
)
85
Yin
 
XM
Oltvai
 
ZN
Korsmeyer
 
SJ
BH1 and BH2 domains of Bcl-2 are required for inhibition of apoptosis and heterodimerization with bax.
Nature
1994
, vol. 
369
 (pg. 
321
-
333
)
86
Chittenden
 
T
Flemington
 
C
Houghton
 
AB
, et al. 
A conserved domain in Bak, distinct from BH1 and BH2, mediates cell death and protein binding functions.
EMBO J
1995
, vol. 
14
 (pg. 
5589
-
5596
)
87
Strasser
 
A
The role of BH3-only proteins in the immune system.
Nat Rev Immunol
2005
, vol. 
5
 (pg. 
189
-
200
)
88
Sattler
 
M
Liang
 
H
Nettesheim
 
D
, et al. 
Structure of Bcl-xL-Bak peptide complex: recognition between regulators of apoptosis.
Science
1997
, vol. 
275
 (pg. 
983
-
986
)
89
Oltersdorf
 
T
Elmore
 
SW
Shoemaker
 
AR
, et al. 
An inhibitor of Bcl-2-family proteins induces regression of solid tumors.
Nature
2005
, vol. 
435
 (pg. 
677
-
681
)
90
Wang
 
JL
Liu
 
D
Zhang
 
ZJ
, et al. 
Structure-based discovery of an organic compound that binds Bcl-2 protein and induces apoptosis of tumor cells.
Proc Natl Acad Sci U S A
2000
, vol. 
97
 (pg. 
7124
-
7129
)
91
Reed
 
JC
Pellecchia
 
M
Apoptosis-based therapies for hematological malignancies.
Blood
2005
, vol. 
106
 (pg. 
408
-
418
)
92
May
 
WS
Tyler
 
PG
Ito
 
T
Armstrong
 
DK
Qatsha
 
KA
Davidson
 
NE
Interleukin- 3 and bryostatin-1 mediate hyperphosphorylation of Bcl-2 α in association with suppression of apoptosis.
J Biol Chem
1994
, vol. 
269
 (pg. 
26865
-
26870
)
93
Itoh
 
T
Deng
 
X
Carr
 
B
May
 
WS
Bcl-2 phosphorylation required for antiapoptosis function.
J Biol Chem
1997
, vol. 
272
 (pg. 
11671
-
11673
)
94
Chen
 
J
Flannery
 
JG
LaVail
 
MM
Steinberg
 
RH
Xu
 
J
Simon
 
MI
bcl-2 overexpression reduces apoptotic photoreceptor cell death in three different retinal degenerations.
Proc Natl Acad Sci U S A
1996
, vol. 
93
 (pg. 
7042
-
7047
)
95
Cheng
 
E
Clem
 
R
Ravi
 
R
, et al. 
Conversion of Bcl-2 to a Bax-like death effector by caspases.
Science
1997
, vol. 
278
 (pg. 
1966
-
1968
)
96
Lin
 
B
Kolluri
 
SK
Lin
 
F
, et al. 
Conversion of Bcl-2 from protector to killer by interaction with nuclear orphan receptor TR3/NGFI-B/Nur77.
Cell
2004
, vol. 
116
 (pg. 
527
-
540
)
97
Bivona
 
TG
Quatela
 
SE
Bodemann
 
BO
, et al. 
PKC regulates a farnesyl electrostatic switch on K-Ras that promotes its association with Bcl-XL on mitochondria and induces apoptosis.
Mol Cell
2006
, vol. 
21
 (pg. 
481
-
493
)
98
Schwartz
 
PS
Manion
 
MK
Emerson
 
CB
, et al. 
2-Methoxy antimycin reveals a unique mechanism for Bcl-x(L) inhibition.
Mol Cancer Ther
2007
, vol. 
6
 (pg. 
2073
-
2080
)
99
Silvestrini
 
R
Veneroni
 
S
Daidone
 
MG
, et al. 
The Bcl-2 protein: a prognostic indicator strongly related to p53 protein in lymph node-negative breast cancer patients.
J Natl Cancer Inst
1994
, vol. 
86
 (pg. 
499
-
504
)
100
Demaurex
 
N
Distelhorst
 
C
Apoptosis–the calcium connection.
Science
2003
, vol. 
300
 (pg. 
65
-
67
)
101
Puthalakath
 
H
Strasser
 
A
Keeping killers on a tight leash: transcriptional and post-translational control of the pro-apoptotic activity of BH3-only proteins.
Cell Death Differ
2002
, vol. 
9
 (pg. 
505
-
512
)
102
Opferman
 
JT
Korsmeyer
 
SJ
Apoptosis in the development and maintenance of the immune system.
Nat Immunol
2003
, vol. 
4
 (pg. 
410
-
415
)
103
Veis
 
DJ
Sorenson
 
CM
Shutter
 
JR
Korsmeyer
 
SJ
Bcl-2-deficient mice demonstrate fulminant lymphoid apoptosis, polycystic kidneys, and hypopigmented hair.
CELL
1993
, vol. 
75
 (pg. 
229
-
240
)
104
Knudson
 
CM
Tung
 
KSK
Tourtellotte
 
WG
Brown
 
GAJ
Korsmeyer
 
SJ
Baxdeficient mice with lymphoid hyperplasia and male germ cell death.
Science
1995
, vol. 
270
 (pg. 
96
-
99
)
105
Ma
 
A
Pena
 
JC
Chang
 
B
, et al. 
Bclx regulates the survival of double-positive thymocytes.
Proc Natl Acad Sci U S A
1995
, vol. 
92
 (pg. 
4763
-
4767
)
106
Zinkel
 
SS
Ong
 
CC
Ferguson
 
DO
, et al. 
Proapoptotic BID is required for myeloid homeostasis and tumor suppression.
Genes Dev
2003
, vol. 
17
 (pg. 
229
-
239
)
107
Bouillet
 
P
Metcalf
 
D
Huang
 
DC
, et al. 
Proapoptotic Bcl-2 relative Bim required for certain apoptotic responses, leukocyte homeostasis, and to preclude autoimmunity.
Science
1999
, vol. 
286
 (pg. 
1735
-
1738
)
108
Bouillet
 
P
Purton
 
JF
Godfrey
 
DI
, et al. 
BH3-only Bcl-2 family member Bim is required for apoptosis of autoreactive thymocytes.
Nature
2002
, vol. 
415
 (pg. 
922
-
926
)
109
Villunger
 
A
Scott
 
C
Bouillet
 
P
Strasser
 
A
Essential role for the BH3-only protein Bim but redundant roles for Bax, Bcl-2, and Bcl-w in the control of granulocyte survival.
Blood
2003
, vol. 
101
 (pg. 
2393
-
2400
)
110
Fischer
 
SF
Bouillet
 
P
O'Donnell
 
K
Light
 
A
Tarlinton
 
DM
Strasser
 
A
Proapoptotic BH3-only protein Bim is essential for developmentally programmed death of germinal center-derived memory B cells and antibody-forming cells.
Blood
2007
, vol. 
110
 (pg. 
3978
-
3984
)
111
Opferman
 
JT
Iwasaki
 
H
Ong
 
CC
, et al. 
Obligate role of anti-apoptotic MCL-1 in the survival of hematopoietic stem cells.
Science
2005
, vol. 
307
 (pg. 
1101
-
1104
)
112
Opferman
 
JT
Letai
 
A
Beard
 
C
Sorcinelli
 
MD
Ong
 
CC
Korsmeyer
 
SJ
Development and maintenance of B and T lymphocytes requires antiapoptotic MCL-1.
Nature
2003
, vol. 
426
 (pg. 
671
-
676
)
113
Takeuchi
 
O
Fisher
 
J
Suh
 
H
Harada
 
H
Malynn
 
BA
Korsmeyer
 
SJ
Essential role of BAX, BAK in B cell homeostasis and prevention of autoimmune disease.
Proc Natl Acad Sci U S A
2005
, vol. 
102
 (pg. 
11272
-
11277
)
114
Print
 
C
Loveland
 
K
Gibson
 
L
, et al. 
Apoptosis regulator bcl-w is essential for spermatogenesis but appears otherwise redundant.
Proc Natl Acad Sci
1998
, vol. 
95
 (pg. 
12424
-
12431
)
115
Ranger
 
AM
Zha
 
J
Harada
 
H
, et al. 
Bad-deficient mice develop diffuse large B cell lymphoma.
Proc Natl Acad Sci U S A
2003
, vol. 
100
 (pg. 
9324
-
9329
)
116
Hamasaki
 
A
Sendo
 
F
Nakayama
 
K
Ishida
 
N
Negishi
 
I
Hatakeyama
 
S
Accelerated neutrophil apoptosis in mice lacking A1-a, a subtype of the bcl-2-related A1 gene.
J Exp Med
1998
, vol. 
188
 (pg. 
1985
-
1992
)
117
Xiang
 
Z
Ahmed
 
AA
Moller
 
C
Nakayama
 
K
Hatakeyama
 
S
Nilsson
 
G
Essential role of the prosurvival bcl-2 homologue A1 in mast cell survival after allergic activation.
J Exp Med
2001
, vol. 
194
 (pg. 
1561
-
1569
)
118
Reed
 
JC
Doctor
 
KS
Godzik
 
A
The domains of apoptosis: a genomics perspective.
Science STKE
2004
, vol. 
2004
 pg. 
RE9
 
119
Luciano
 
F
Krajewska
 
M
Ortiz-Rubio
 
P
, et al. 
Nur77 converts phenotype of Bcl-B, an anti-apoptotic protein expressed in plasma cells and myeloma.
Blood
2007
, vol. 
109
 (pg. 
3849
-
3855
)
120
Russell
 
HR
Lee
 
Y
Miller
 
HL
Zhao
 
J
McKinnon
 
PJ
Murine ovarian development is not affected by inactivation of the Bcl-2 family member diva.
Mol Cell Biol
2002
, vol. 
22
 (pg. 
6866
-
6870
)
121
O'Brien
 
S
Moore
 
JO
Boyd
 
TE
, et al. 
Randomized phase III trial of fludarabine plus cyclophosphamide with or without oblimersen sodium (Bcl-2 antisense) in patients with relapsed or refractory chronic lymphocytic leukemia.
J Clin Oncol
2007
, vol. 
25
 (pg. 
1114
-
1120
)
122
Han
 
YH
Cao
 
X
Lin
 
B
, et al. 
Regulation of Nur77 nuclear export by c-Jun Nterminal kinase and Akt.
Oncogene
2006
, vol. 
25
 (pg. 
2974
-
2986
)
Sign in via your Institution