GENETIC CHANGES involving oncogenes and tumor suppressor genes contribute to the deregulated expansion of malignant cells. While some of these changes result in increased proliferation, others contribute to an increase in cell numbers by inhibiting apoptosis (programmed cell death).1 Because cytotoxic drugs or irradiation result in cell killing by apoptosis, the genetic changes underlying malignancy often reduce the ability of these agents to destroy malignant cells.1 2 The elucidation of the pathways involved in the regulation of apoptosis in normal and malignant hematopoietic cells is therefore likely to contribute to the development of improved therepeutic stategies in the treatment of leukemia and lymphoma. This review first summarizes recent advances in the understanding of the control of apoptosis. Examples of how this control is altered in leukemic cells is then described.

Apoptosis is a tightly regulated form of physiological cell death which is dependent on the expression of cell-intrinsic suicide machinery.3 Prominent morphological changes include cell shrinkage, condensation of the nuclear chromatin, fragmentation of the nucleus, and cleavage of chromosomal DNA at internucleosomal sites, resulting in the generation of a characteristic ladder pattern of DNA fragments on electrophoresis. Blebbing of the cell surface results in the release of membrane-bound apoptotic bodies.3Phosphatidylserine, which is normally located on the inner face of the plasma membrane, becomes exposed on the outer surface and provides a recognition signal for engulfment by phagocytes.4 5 Thus, apoptosis results in the rapid and efficient removal of superfluous or damaged cells.

Genetic studies in the nematode C elegans have resulted in the identification of a set of genes involved in the regulation of apoptosis.6 The ced-3 gene encodes a cysteine protease, which is homologous to members of the caspase protease family that execute the apoptotic program in mammalian cells (see section 3.1). The ced-4 gene product is required for the activation of CED-3. This activation step is blocked by the CED-9 protein, which is homologous to mammalian BCL-2. BCL-2 can substitute for CED-9 in blocking apoptosis in C elegans7,8 whereas overexpression of CED-4 induces apoptosis in mammalian cells,9 suggesting a high degree of conservation of the mechanisms of apoptosis regulation. Therefore, the C elegansmodel has been of value in the identification of the proteins that control apoptosis in human cells (see sections 3 and 4).

The induction of apoptosis may conveniently be divided into three stages: (1) the interaction of the inducing signal with the cell, (2) biochemical transduction of the death signal, and (3) the execution of apoptosis. Because different extracellular signals and signal transduction pathways converge on a final common pathway during the execution phase, this terminal stage of apoptosis will be summarized first.

3.1. The caspase family of proteases mediates the terminal stages of apoptotic cell death.

The terminal stages of apoptosis involve the activation of a related family of proteases, the caspases.10,11 These enzymes possess an essential cysteine residue within their active sites and cleave substrates adjacent to aspartate residues. The cDNAs encoding 10 caspases have been cloned.10 

Caspases are expressed as inactive pro-enzymes. Cleavage of these pro-caspases adjacent to aspartate residues results in the generation of active subunits of approximately 10 and 20 kD (Fig 1). These subunits dimerize, with the resultant generation of a complete active site.10 The N-terminal pro-domains (Fig 1) of some pro-caspases, which are removed during activation, nevertheless play important roles in mediating regulatory interactions between caspases and other proteins. The requirement for cleavage adjacent to aspartate residues during caspase activation, together with the aspartate specificity of caspases, raises the possibility that cascades of caspase activation events may be involved in apoptosis regulation (see sections 3.3 and 4.3).

Fig. 1.

Structure and activation of caspases. The peptide sequences that contribute to the enzymatically active caspase are shown in cross-hatch. Asp, aspartate residues at which the inactive pro-caspase is proteolytically cleaved with the resultant generation of the active subunits.

Fig. 1.

Structure and activation of caspases. The peptide sequences that contribute to the enzymatically active caspase are shown in cross-hatch. Asp, aspartate residues at which the inactive pro-caspase is proteolytically cleaved with the resultant generation of the active subunits.

Close modal

Caspases 3, 6, and 7 are terminal members of caspase cascades and recognize critical cellular substrates, whose cleavage contributes to the morphological and functional changes associated with apoptosis.10 Caspase 3 substrates include poly (ADP-ribose) polymerase,12,13 an enzyme involved in regulation of DNA repair and gelsolin, a cytoskeletal protein.14 Caspase 3 activation also results in DNA cleavage via inactivation of an inhibitor of DNA fragmentation factor, the endonuclease responsible for internucleosomal cleavage of chromatin.15 Caspase 6 substrates include the nuclear structural protein, lamin.16Thus, the cleavage of a relatively restricted set of critical caspase substrates contributes to the apoptotic demise of cells via disassembly of structural components, cleavage of the genetic material, and prevention of DNA repair.

3.2. Specific protease inhibitors block cell death by targetting terminal caspases.

In vitro studies suggest that the inhibitor of apoptosis (IAP) family of proteins may modulate cell death via abrogation of caspase activity.

These proteins are similar to the baculovirus-encoded caspase inhibitor p35. IAP1 and 2, XIAP (X-linked IAP) and survivin contain one to three BIR (baculovirus IAP repeat) motifs that are essential to their function.17 IAPs 1 and 2 and XIAP specifically target caspases 3 and 7, which function at the distal end of proteolytic cascades. Therefore, IAP expression may serve to reprieve cells otherwise committed to apoptotic death. However, the Drosophila IAPs (DIAP-1 and DIAP-2) block apoptosis by direct binding via the BIR motifs to noncaspase death-inducing proteins encoded by thereaper, hid, and grim genes.18Therefore, it is possible that the mammalian IAPs may also inhibit cell death by mechanisms other than binding to caspases.

The key question regarding cell death regulation concerns the mechanisms by which caspase activation steps are triggered by apoptotic signals. Different extracellular signals interact with the caspase system in different ways.

3.3. Ligation of FAS or the tumor necrosis factor (TNF) receptor results in the direct activation of caspases.

FAS and the TNF receptor are structurally related transmembrane receptor proteins. Their extracellular domains bind FAS ligand and TNF, respectively, resulting in the formation of receptor trimers. The cytoplasmic domain of FAS contains a “death domain” (Fig 2), whose elimination results in the abrogation of cell killing.19 The death domain of FAS recruits the death domain of the FADD (Fas-associated death domain) protein following receptor trimerization.20 FADD also contains a death-effector domain, which mediates interaction with similar amino acid sequences in the pro-domain of pro-caspase 8.21 22 Trimerization of the FAS/FADD/pro-caspase 8 complex after ligand binding results in the cleavage of the pro-caspase and generation of active caspase 8. Caspase 8 then cleaves pro-caspase 3, probably via activation of an unidentified intermediate caspase (Fig 2).

Fig. 2.

Signal transduction by FAS. Protein-protein interactions between the death domains (horizontal stripes) of FAS and FADD and the death-effector domains (stippled) of FADD and pro-caspase 8 are shown. Trimerization of receptor complexes after binding of FAS to FAS ligand results in the cleavage of pro-caspase 8 and the release of active caspase 8.

Fig. 2.

Signal transduction by FAS. Protein-protein interactions between the death domains (horizontal stripes) of FAS and FADD and the death-effector domains (stippled) of FADD and pro-caspase 8 are shown. Trimerization of receptor complexes after binding of FAS to FAS ligand results in the cleavage of pro-caspase 8 and the release of active caspase 8.

Close modal

Activation of apoptosis after TNF receptor ligation follows a similar pattern. However, the TNF receptor does not bind FADD directly, but does so via a linking protein, TRADD (TNF receptor-associated death domain).19 

3.4. Apoptosis induction by the perforin/granzyme system.

Killing of target cells by cytotoxic T lymphocytes plays a major role in defense against malignant and virus-infected cells, and contributes to transplant rejection and autoimmune disease. Killing is preceded by the release of the contents of cytotoxic T-cell granules, which contain perforin and the serine proteases granzymes A and B. Perforin forms a pore in the plasma membrane of the target cell, thereby allowing entry of granzyme B into the cytosol.23 Granzyme B cleaves and activates caspase 3 in cell-free systems.24,25 However, the primary target of granzyme B in intact cells is likely to be caspase 10, whose activation results in the subsequent activation of caspase 3.26 In cell-free systems, the addition of granzyme B initiates cleavage of several apoptosis-specific substrates and also induces chromatin condensation. Abrogation of these events by selective inhibitors suggests that activation of caspase 3 (and possibly of caspase 7) may be important mediators of apoptosis induction by cytotoxic T-cell–derived granzyme B.26 However, genetic studies have shown that the granules of cytotoxic T cells contain additional cytotoxic components in addition to perforin and granzymes A and B.27 

4.1. The BCL-2 protein family plays a central role in the regulation of apoptosis.

The 26-kD BCL-2 protein protects cells from the induction of apoptosis by diverse stimuli, including the withdrawal of survival factors, heat shock, and treatment with DNA damaging agents.28-30 BCL-2 is the prototype of a family of related proteins. Other anti-apoptotic family members include BCL-XL, BCL-w, MCL-1, and A1. In contrast, the BAX, BAK, and BAD proteins are examples of pro-apoptotic BCL-2 family members whose overexpression promotes cell killing.30 The conserved BH1 (BCL-2 homology 1) and BH2 domains of the anti-apoptotic proteins form a hydrophobic cleft which binds the BH3 domains of pro-apoptotic family members, at least in vitro.31 32 

The susceptibility of cells to apoptosis is determined in part by the relative concentrations of pro- and anti-apoptotic BCL-2 family members. The antagonistic actions of these two groups of proteins have been attributed to their ability to form heterodimers.33However, at least some of the dimerization properties of the BCL-2 family may be artefacts induced by detergents in vitro.34Furthermore, genetic studies suggest that BCL-2 and BAX function independently of one another in the regulation of apoptosis.35 Deletion of the BH4 domain of BCL-2 impairs its ability to block apoptosis without affecting its dimerization with anti-apoptotic family members.9 Deletion of the BH3 domain of BAX, which is required for its dimerization, does not impair the ability of the protein to increase the sensitivity of cells to cytotoxic agents.36 Therefore, it is plausible that the actions of pro- and anti-apoptotic members of the BCL-2 family may determine the sensitivity of cells to apoptosis induction via binding to a common target rather than to dimer formation.35 36 

BCL-2 targets to the outer mitochondrial membrane, the nuclear envelope, and the endoplasmic reticulum via its C-terminal hydrophobic domain.30,37,38 However, some studies suggest that BAX shows a largely diffuse sub-cellular localization, translocating rapidly to the mitochondria (and possibly other organelles) after the induction of an apoptotic signal.39 

The BCL-2 family regulates apoptosis induction via control of the activation of caspases, apparently by a mechanism involving the release of mitochondrial cytochrome c.11 However, it is unclear whether cytochrome c release is a component of the primary apoptosis induction pathway or a means of amplifying the death signal, as summarized later (see sections 4.3 and 4.4). The mechanism of cytochrome c–dependent caspase activation is discussed next.

4.2. Cytochrome c triggers caspase 3 cleavage via activation of caspase 9.

Three proteins have been purified from the cytosol of HeLa cells which, when recombined in the presence of adenosine triphosphate (ATP) (or deoxy ATP), were necessary and sufficient for the cleavage and activation of pro-caspase 3. These proteins, originally designated as Apaf 1, 2, and 3 (Apaf = apoptoticprotease activating factor) have been characterized.40,41 Apaf 1 contains a central domain with homology to C elegans CED-4.40 The amino-terminal domain of Apaf 1 is homologous to the CARDs (caspaserecruitment domains) of some caspases (Fig 3). The carboxy terminus consists of several WD repeats, which mediate interactions between certain regulatory proteins. Apaf 2 was found to be identical to cytochrome c,40 while Apaf 3 is identical to caspase 9.41 

Fig. 3.

Apaf-1– and cytochrome c–dependent activation of caspases.

Fig. 3.

Apaf-1– and cytochrome c–dependent activation of caspases.

Close modal

A model11,41,42 for the cytochrome c–dependent activation of caspase 3 is depicted in Fig 3. In the absence of cytochrome c, the WD repeat domain prevents interaction of Apaf-1 with pro-caspase 9. Cytochrome c binds Apaf 1, inducing a conformational change that results in the interaction of Apaf 1 and caspase 9, mediated by the CARD pro-domains present on both these proteins. Apaf-1 induces the cleavage and activation of pro-caspase 9 via a mechanism involving oligomerization of Apaf-1, thus facilitating autocatalytic cleavage of the pro-caspase.42 Activated caspase 9 now cleaves and activates caspase 3.

4.3. The BCL-2 protein family apparently regulates the release of cytochrome c from mitochondria.

Cytochrome c is released from mitochondria during apoptosis induced by diverse stimuli.11 Overexpression of BCL-2 or BCL-XL inhibit cytochrome c release induced by etoposide, actinomycin D, oxidative stress, Fas ligation, or interleukin-3 (IL-3) withdrawal.43-45 The BAX protein, on the other hand, triggers redistribution of cytochrome c in the absence of apoptotic stimuli.46 Thus, BCL-2 and BCL-XL may prevent apoposis by inhibiting cytochrome c release while BAX favors cell death by promoting its relocation to the cytosol. However, it is unclear whether these actions of the BCL-2 family result from the direct actions of these proteins on mitochondria, which then initiate caspase activation or whether cytochrome c release is secondary to BCL-2 family-regulated caspase activation and plays a subsequent role in amplification of the apoptotic signal.47 

Some evidence suggests a direct role for the BCL-2 family in regulating cytochrome c release. The structure of BCL-XL resembles that of pore-forming bacterial toxins.31 BCL-2, BCL-XL, and BAX form ion channels in vitro.48-50 Because pores formed by BAX protein may show high conductance values under some conditions,50 it is possible that these channels allow the exit of cytochrome c. Anti-apoptotic proteins including BCL-2 may block release by interfering with pore formation by BAX.50 However, channel formation by BCL-2 family proteins has only been shown in synthetic membranes and, in some cases, at nonphysiological pH.48-50Therefore, it has not been established that these proteins can form channels in cellular membranes under physiological conditions.

An alternative hypothesis suggests that BCL-2 family proteins regulate the electrical potential gradient (Δψm) across the inner mitochondrial membrane and thereby regulate mitochodrial volume. The opening of “megachannels” in the inner membrane allows the passage of molecules of less than 1.5 kD, resulting in the dissipation of Δψm. Apoptosis induced by stimuli including antineoplastic drugs and glucocorticoids is apparently preceded by disruption of the gradient. Both the loss of Δψm and subsequent apoptosis induction are blocked by the megachannel antagonist bongkrekic acid or by overexpression of BCL-2 or BCL-XL.51,52 Therefore, changes in the permeability of the inner mitochondrial membrane may be a central coordinating event in the induction of apoptosis, which is inhibited by anti-apoptotic BCL-2 family members.51 The reported ability of BCL-2 to modulate ion fluxes across the inner mitochondrial membrane53 is compatible with this hypothesis. The precise mechanistic relationship between loss of Δψm and cytochrome c release has not been established. However, because loss of Δψm results in mitochondrial swelling, Δψm loss may cause cytochrome c release as a result of outer membrane rupture.45 

It is now apparent that the BCL-2 family regulates steps in apoptotic signaling distal to cytochrome c release. BCL-2 overexpression blocks apoptosis in cells containing high concentrations of cytosolic cytochrome c, induced either by transient expression of BAX54 or by direct microinjection.55 These observations may be accounted for by the direct action of some anti-apoptotic BCL-2 family members on caspase activation, because BCL-XL can bind Apaf-1 directly and block its ability to activate pro-caspase 9.56 Such a function for BCL-2 family proteins parallels the role of the C elegans BCL-2 homologue CED-9.57 A model that combines the putative ability of BCL-2 family members to regulate both cytochrome c release and the Apaf-1–mediated activation of pro-caspase 9 (the “Swiss army knife” model47) is depicted in Fig 4A.

Fig. 4.

Models relating the roles of BCL-2 family proteins, cytochrome c release, and Apaf-1–dependent caspase activation during the induction of apoptosis. (A) “Swiss army knife” model; (B) “death cycle” model. See text for details.

Fig. 4.

Models relating the roles of BCL-2 family proteins, cytochrome c release, and Apaf-1–dependent caspase activation during the induction of apoptosis. (A) “Swiss army knife” model; (B) “death cycle” model. See text for details.

Close modal
4.4. Caspase action on mitochondria amplifies the initial apoptotic signal via a positive feedback loop.

Caspases can themselves trigger cytochrome c release, because selective inhibitors of these proteases can abrogate release in response to some stimuli.45 Furthermore, death signals including ligation of Fas, which directly activate caspases (section 4.2), may nevertheless be amplified via caspase-mediated cytochrome c release.45Recombinant caspases disrupt Δψm and induce cytochrome c release when added to isolated mitochondria.58 The mechanism of this action of caspases is unclear.

The operation of a positive feedback loop raises the possibility that anti-apoptotic BCL-2 family members may not play a direct role in the modulation of cytochrome c release. The ability of these proteins to directly inhibit Apaf-1 function56 could instead modulate cytochrome c release indirectly via the caspase-dependent feedback loop. This “death cycle” model47 implies that cytochrome c release is not a component of the apoptosis-initiating pathway but serves in the amplification of an initial signal generated via the regulation of Apaf-1 (Fig 4B).

4.5. Survival factors regulate apoptosis via phosphorylation of the BAD protein.

The BAD protein is a pro-apoptotic member of the BCL-2 family.30,59 FL5.12 lymphoid cells depend on IL-3 for their survival in vitro. In cells cultured in the presence of IL-3, the BAD protein is phosphorylated on serine residues. Phosphorylated BAD is sequestered via binding to the 14-3-3 protein and is, therefore, unable to promote apoptosis.59 The BAD protein rapidly becomes dephosphorylated in the absence of IL-3, dissociates from 14-3-3, and triggers apoptosis (Fig 5).

Fig. 5.

Apoptosis modulation by IL-3 via phosphorylation of the pro-apoptotic BAD protein.

Fig. 5.

Apoptosis modulation by IL-3 via phosphorylation of the pro-apoptotic BAD protein.

Close modal

Recent evidence has implicated protein kinase B as the protein kinase that mediates BAD phosphorylation in response to IL-360 and other survival factors61 (Fig 4). IL-3 triggers activation of phosphatidylinositol 3-kinase, a lipid kinase that initiates the eventual generation of phosphatidylinositol 3,4-bisphosphate, an allosteric activator of protein kinase B.62 

The p53 protein plays an important role in the coupling of DNA damage to cell-cycle arrest and to the induction of apoptosis. In cells with undamaged DNA, p53 protein levels are maintained at a low level as a result of rapid turnover. An increase in stability after the induction of DNA damage results in an increased level of p53.63 The protein product of the ataxia telangiectasia (atm) gene participates in a pathway that links the detection of DNA damage to the upregulation of p53.64 However, the carboxy terminus of the p53 protein itself can bind to damaged DNA,65 suggesting that both p53 and the putative damage detector may colocalize at the site of DNA damage. The radiation resistance of thymocytes derived from p53 “knockout” mice when compared with wild-type thymocytes66,67 emphasizes the importance of p53-dependent mechanisms in the induction of apoptosis after DNA damage induction. Upregulation of p53 also results in cell-cycle arrest. The pathways involved in this facet of p53 action have recently been reviewed.63 

5.1. Transcriptional activation by p53.

A tetramer of p53 molecules functions as a transcription factor that binds to consensus sequences in the 5′ untranslated regions of specific target genes.63 The upstream region of the BAX gene contains p53 consensus binding sites.68 Enforced p53 expression augments BAX expression, which is followed by apoptosis induction.69 Genetic studies on apoptosis induction in adriamycin-treated mouse fibroblasts suggest that BAX is an important (but not the only) effector of p53-mediated apoptosis.70However, thymocytes isolated from p53 “knockout” mice expressing elevated levels of BAX are as resistant to etoposide-induced apoptosis as are thymocytes from p53 “knockout” mice expressing normal levels of bax.71 Therefore, mechanisms other than BAX induction may mediate p53-dependent apoptosis, at least in some cell types.

Polyak et al72 have identified 12 mRNA species which were rapidly induced after adenovirus-mediated transfer of the p53 gene into p53 null colorectal carcinoma cells. Thesep53-induced genes (PIGs) either encode proteins that catalyze redox reactions and consequently generate reactive oxygen species (ROS) or whose expression is augmented by ROS generation. Indeed, introduction of p53 into p53 null cells results in a burst of ROS generation that is followed by apoptosis induction. Inhibitors of ROS generation do not interfere with induction of PIG genes by p53 but do abrogate apoptosis induction, suggesting that the upregulated expression of PIG genes and the subsequent generation of ROS play a critical role in the induction of apoptosis by p53 at least in some cell types.72 ROS can themselves trigger cytochrome c release from mitochondria,44 possibly via modulation of ion transport within these organelles,53 73 suggesting an additional mechanism for the induction of apoptosis by p53.

Transient expression of p53 results in ROS generation in cells that are susceptible to p53-mediated apoptosis but not in resistant cells, compatible with the hypothesis that ROS are downstream mediators of p53-induced apoptosis.74 However, the role of ROS in the regulation of apoptosis remains controversial, because in some studies the induction of cell death is not abrogated at very low oxygen tension.75 76 

5.2. Transcriptional repression by p53.

In addition to its transactivating properties, p53 represses transcription from several promoters that lack p53 binding sites. BCL-2 can relieve this transcriptional repression and also protect cells from apoptosis, suggesting that inhibition of transcription of specific but as yet unidentified genes may contribute to the ability of p53 to induce apoptosis.77 However, p53 is also able to induce apoptosis via pathways that are not dependent on the regulation of gene expression.78 Therefore, induction of apoptosis by p53 can occur by diverse pathways depending on the cellular context. It is also clear that some apoptotic pathways do not involve p53. For example, thymocytes from p53 knockout mice are resistant to etoposide and radiation but not to glucocorticoids.67 Furthermore, HL60 cells, which have lost both p53 alleles, are extremely sensitive to apoptosis induction by drugs that induce DNA strand-breaks.79 The p53 dependence of apoptotic pathways is also tissue dependent, because radiation-induced apoptosis is compromised in the thymus of p53 knockout mice, but not in the lung.80 

5.3. p53 loss results in resistance to cytotoxic regimes.

When transplanted into imunodeficient mice, fibrosarcomas expressing functional p53 show a high proportion of apoptotic cells and regress after treatment with adriamycin or γ radiation. In contrast, tumors lacking p53 show few apoptoses and are resistant to adriamyin or radiation.81 Therefore, inactivation of p53 can result in the resistance of tumors to DNA damaging agents. The elevation of BAX expression in response to radiation is only detected in human leukemia cell lines that express p53 and that die by apoptosis in response to DNA damage. By contrast, p53-negative lines do not elevate BAX levels and are also resistant to radiation-induced apoptosis.82Therefore, p53-mediated elevation of BAX contributes to the killing of some tumor cells by cytotoxic regimes.77 

5.4. The p53-related p73 gene product.

The p73 gene encodes a protein that is closely related to p53.83 Overexpresion of p73 results in the induction of some genes that are also targets of p53, and also induces apoptosis. However, p73 expression is apparently not augmented after the induction of DNA damage.83 Therefore, there is at present no evidence implicating p73 in DNA damage-induced cell killing.

During hematopoiesis, the survival of progenitor cells is regulated both positively and negatively by a complex, interacting network of cytokines and adhesion molecules.84 Noncycling primitive CD34+ human hematopoietic progenitors require the continuous presence of IL-3 or granulocyte-macrophage colony-stimulating factor (GM-CSF) for survival in vitro. In contrast, other cytokines including IL-6 and IL-11 trigger proliferation of these progenitors.85 Stem cell factor, Flt ligand, and IL-3 suppress apoptosis in single-cell assays designed to test the direct actions of cytokines on primitive progenitors. Thrombopoietin is more effective in preventing apoptosis than any of these cytokines.86 Cytokines show target cell selectivity in preventing apoptosis. For example, stem cell factor selectively promotes survival of primitive hematopoietic cells, whereas IL-3 blocks cell death in more committed progenitors.87 Flt ligand is selective for progenitors committed to the myeloid lineage.88 

Other cytokines promote the apoptotic death of both primitive and committed progenitors.84 The flt3 ligand-mediated survival of primitive progenitors is counteracted by both transforming growth factor-β (TGF-β) and TNF-α.89 Interferon-γ (IFN-γ) suppresses the survival of long-term culture-initiating cells. The action of IFN-γ is more potent when this cytokine is secreted by stromal cells in culture than when added to the medium, stressing the importance of the hematopoietic microenvironement in modulating survival.90 Induction of apoptosis by both IFN-γ and TNF-α may be mediated in part by increasing expression of FAS on the surface of hematopoietic progenitors.91Subsequent ligation of this death receptor then triggers cell killing.

Primitive bone marrow B-lymphoid progenitors require direct contact with bone marrow stromal cells for survival.92 These survival-promoting interactions are dependent on interactions between the β1 integrins VLA-4 and VLA-5 expressed on the B-cell surface and fibronectin generated by the fibroblasts.93 It is unclear whether interactions between these adhesion molecules directly generate survival signals or whether the close juxtaposition of the lymphoid progenitors to fibroblasts enhances the actions of unidentified survival factor generated by the fibroblasts.

Apoptotic death of progenitor cells following deprivation of survival factors is an active rather than a passive process. Hematopoietic cells from p53 “knockout” mice are more resistant to the induction of apoptosis after factor withdrawal than are corresponding cells from control animals.94 Murine 32Dc13 myeloid precursor cells depend on IL-3 for survival in vitro. Withdrawal of IL-3 results in apoptotic death, which is dependent on the expression of wild-type p53,95 suggesting that the activation of a cell-intrinsic pathway involving p53 is a prerequisite for cell killing after removal of survival factors.

Cytokines modulate both the basal survival of some leukemia cell lines and also compromise their killing by cytotoxic treatments.84 G-CSF, GM-CSF, IL-3, IL-6, or IFN-γ protect murine myeloid leukemia cell lines from apoptotic death induced by cytotoxic drugs.96,97 Apoptosis induced by the introduction of wild-type p53 into a p53-negative murine AML cell line is abrogated by IL-6,98 suggesting that cell killing on deprivation of this cytokine proceeds via a p53-mediated pathway.

The generation of the recognition repertoires of T and B lymphocytes is dependent on the apoptotic deletion of cells with inappropriate specificities.99 Killing of cells after ligation of FAS or the receptor proceeds via induction of apoptosis.19 Activation of T lymphocytes after encounters with cognate antigen/major histocompatibility complexes (MHC) results in activation and concomitant upregulation of FAS ligand. Subsequent interactions between FAS ligand and FAS induces apoptotic death of the activated T cells, thereby downregulating the immune response. Elimination of autoreactive B lymphocytes is also mediated by the FAS system.19 Triggerring of specific cytotoxic T cells by viral antigens displayed at the surface of target cells induces expression of FAS ligand. Interaction of the ligand with FAS expressed by the target cell initiates apoptosis.19 Cytotoxic T lymphocytes also kill target cells via the perforin/granzyme system (section 3.4).

Treatment of leukemia cell lines with cytotoxic drugs results in the release of cytochrome c43-45 and the activation of caspases.100,101 Caspases are also activated after cytotoxic treatment of freshly isolated B chronic lymphocytic leukaemia (B-CLL) cells.102 However, the mechanisms that couple DNA damage to more downstream regulatory events are largely unclear. Evidence of a largely circumstantial nature suggests that some of the mechanisms of apoptosis control described in sections 3, 4, and 5 are deregulated in leukemia cells, thus contributing to their abnormal expansion and, in some cases, to drug and radiation resistance. Deregulation of apoptosis results from translocations involving genes that encode cell death–regulating proteins. However, microenvironmental factors also impinge on both the basal survival of leukemia cells and their killing by cytotoxic regimes. Some of this evidence is summarized next.

8.1. Translocation of the BCL-2 gene in non-Hodgkin’s lymphoma (NHL).

The t(14;18) chromosomal translocation associated with NHL results in the juxtaposition of the BCL-2 gene to the Ig heavy chain (IgH) locus.103 Translocation results in enhanced levels of BCL-2 mRNA, which may be partially attributable to the presence of a powerful transcriptional enhancer in the IgH locus.98 The efficiency of splicing of BCL-2 exons is also increased as a result of their fusion to Ig gene introns in t(14;18) cells. The resulting increase in cellular levels of spliced BCL-2 open reading frames also contributes to the upregulation of BCL-2 protein levels in NHL cells.104 Transgenic mice carrying a BCL-2 gene expressed via the IgH gene enhancer overexpress BCL-2 specifically in B-lymphoid cells. These mice accumulate abnormal numbers of small, nonproliferating B cells that show extended survival in vitro.105,106 BCL-2 overexpression alone is insufficient for lymphomagenesis. However, doubly transgenic mice in which overexpression of both BCL-2 and c-MYC is targetted to B-lymphoid cells rapidly develop tumors originating from primitive lymphoid-committed lymphoid cells, suggesting that a second genetic event is necessary for the malignant transformation of lymphocytes overexpressing BCL-2.106 

Enforced overexpression of the BCL-2 or BCL-XL gene in leukemia cell lines confers increased resistance to cytotoxic drugs.107-109 However, low-grade NHL patients with increased BCL-2 expression respond well to chemotherapy, although complete remissions are rare.110 Although it is likely that overexpression of BCL-2 impairs apoptosis induction in follicular lymphoma cells, additional microenvironemental signals are required to maintain cell viability. NHL cells remain viable for 1 to 2 days in culture and then die rapidly.111 Cell death is preceded by the downregulation of BCL-XL expression, although BCL-2 expression is maintained. Both the decrease in BCL-XL and cell death are prevented by ligation of CD40, suggesting that continuous signaling by this cell-surface molecule is required to maintain viability of the lymphoma cells via upregulation of BCL-XL expression.111 BCL-2 antisense oligodeoxynucleotides have been used in the treatment of nine patients with relapsed NHL. A reduction in tumor mass was observed in two patients and a decrease in circulating tumor cells in two others. In two of five samples that were studied by flow cytometry, a decrease in BCL-2 protein levels was detected after treatment.112 

NHL cells express variable levels of cell-surface FAS, but are resistant to killing after FAS ligation. Therefore, loss of sensitivity to this apoptotic pathway may contribute to the expansion of the lymphoma cells by allowing their escape from normal immune regulatory mechanisms.113 

8.2. BCL-2 expression in CLL.

CLL cells show an extended life span in vivo. They proliferate very slowly, suggesting that a failure to die by apoptosis contributes to the accumulation of malignant cells in this disease.114Translocations of the BCL-2 gene to Ig loci are detected in less than 2% of CLL cases.115 Nevertheless, CLL cells from some patients express high levels of BCL-2 protein compared with BAX.116,117 In vitro, malignant cells isolated from 30% of CLL cases survive for several weeks in the absence of added cytokines.118 However, malignant cells from the remaining 70% of patients undergo rapid apoptosis in culture, but are protected by the addition of cytokines including IL-4 and IFN-α or -γ.118-122 IL-4 and IFN-α may promote CLL cell survival by preventing loss of expression of BCL-2 in vitro.118-120Interaction with bone marrow stroma, which is mediated by the β1 and β2 integrins, also maintains viability of CLL cells.123 124 Normal B cells do not adhere to stroma and are not protected from apoptosis. Therefore, upregulation of integrins on the surface of CLL cells relative to normal B cells may contribute to their extended life span in vivo, via the activation of unknown intracellular pathways.

The ratio of BCL-2 to BAX correlates inversely with the sensitivity of B-CLL cells to cytotoxic drugs in vitro.116,117 However, in vitro sensitivity to fludarabine failed to correlate with the achievement of clinical response.125 In a limited study of 58 CLL patients, high levels of the anti-apoptotic BCL-2 family member MCL-1 were found to correlate significantly with a failure to achieve complete remission.125 

8.3. BCL-2 expression by acute myeloid leukemia (AML) and acute lymphoblastic leukemia (ALL) cells.

Genetic changes that directly result in augmented expression of BCL-2 have not been described in AML. However, those AML patients in which greater than 20% of blasts express detectable BCL-2 levels show shorter survival and lower rates of achievement of complete remission compared with patients whose malignant cells express low BCL-2 levels.126 Immunohistochemical staining of BCL-2 is more intense in malignant cells from AML patients who fail to achieve remission than in those who respond to chemotherapy.127Malignant cells from patients showing high BCL-2 expression are drug resistant in vitro.128 Therefore, high BCL-2 expression, resulting from unknown mechanisms, may confer drug resistance on AML cells. However, other mechanisms may also confer protection even in the absence of high BCL-2 expression, because some AML isolates with low BCL-2 expression are also drug-resistant in vitro.128Elevated expression of MCL-1 at relapse suggests that cytotoxic regimes may result in the selection of AML clones expressing high levels of this anti-apoptotic BCL-2 family protein.129 

Incubation of AML blasts with antisense oligonucleotides designed to decrease BCL-2 expression increases their sensitivity to cytosine arabinoside in vitro,130 emphasizing the potential importance of this protein in conferring drug resistance to these cells.

The expression of BCL-2 by the malignant cells of ALL patients at presentation is highly variable. However, BCL-2 expression does not correlate either with the ability of the ALL cells to survive in vitro or with the response of the patients to intensive chemotherapy.131 Therefore, it is likely that other factors play important roles in modulating the survival of ALL cells. Interactions between B-lineage ALL cells and stromal fibroblasts, mediated by interactions between β1 integrins and fibronectin, promote the survival of the leukemia cells.92 93 The intracellular pathways of survival promotion triggerred by these interactions are, however, unknown.

Deletion and/or mutation of p53 alleles results in the generation of tumors with impaired expression of functional p53 protein.63 The distribution of p53 mutations in human leukemia, which has been extensively reviewed 132, will be briefly summarized here. In general, p53 alterations are more frequent in aggressive disease and are associated with drug resistance and poor survival.

9.1. p53 mutations in myelodysplastic syndromes, CML, and CLL.

p53 changes are seen in 4% of myelodysplastic syndromes and are more frequent in advanced stages.132 p53 mutations are rare in the chronic phase of CML but are more frequent in blast crisis.133 p53 mutations are detected in 10% to 15% of CLL and are associated with poor response to therapy and shorter survival.134-136 Mutations are more frequent (about 40%) in Richter’s immunoblastic transformation.137 However, sensitivity of B-CLL cells to camptothecin analogs or fludarabine in vitro did not correlate with the presence of p53 mutations.116 

9.2. p53 mutations in lymphoma and ALL.

A high proportion (30%) of Burkitt’s lymphoma and 55% of its leukemic counterpart, L3 ALL, harbor p53 mutations in addition to translocation and overexpression of the MYC oncogene.137The transformation of follicular lymphoma to diffuse aggressive disease correlates with p53 mutation and decreased survival in 25% to 30% of cases.138 p53 mutation is associated with loss of the short arm of chromosome 17, which carries the p53 gene, in Ph1-positive ALL.139 Overall, p53 loss is observed in 13% of ALL.132 However, the incidence of these changes is much lower in pediatric ALL (2%) and may correlate with the excellent response of these childhood ALL to cytotoxic drugs.140 

9.3 p53 mutations in AML.

Genetic changes involving p53 are rare in AML, but are more frequent in cases with deletion of chromosome 17p.141 Again, loss of functional p53 in AML is associated with low rates of complete remission and with decreased survival.135 The blast cells from the majority AML patients require the addition of exogenous cytokines for both survival and growth in vitro.142,143Factor-dependent blasts die rapidly when deprived of GM-CSF, but are protected from apoptosis by antisense oligonucleotides that downregulate p53 expression, suggesting that the killing of these cells after factor deprivation is p53-dependent.143 GM-CSF and IL-3 also protect blast cells from 70% of AML patients from apoptosis induction by doxorubicin.144 

In summary, p53 loss is relatively rare in leukemia. However, small sub-populations of leukemia cells may harbor these genetic changes, resulting in their relative resistance to cytotoxic regimes. p53-negative sub-populations may, therefore, survive drug treatment and initiate relapsed disease showing a more aggressive phenotype and increased drug resistance.

9.4. Adenoviruses lacking the E1B gene selectively kill tumor cells lacking functional p53.

Adenovirus infection of human cells requires expression of the viral E1B gene. The product of this gene binds to cell-encoded p53, thereby permitting viral replication and eventual killing of the host cells. Mutant adenoviruses lacking the E1B gene are, therefore, unable to proliferate in normal human cells, but are able to do so in tumor cells lacking functional p53.145 Human cervical carcinomas carried as xenografts in immunodeficient mice regress following direct injection of E1B-negative adenovirus. Primary infection of only 2% of the tumor cells may be sufficient to induce regression, due to the infectious nature of the adenovirus.145 It remains to be established whether the strategy outlined above will result in the selective killing of human leukemia cells.

9.5. Genetic changes involving the atm gene.

Ataxia telangiectasia patients show a high incidence of lymphoid, but not of myeloid, malignancies.146 Sixty percent of patients with T-prolymphocytic leukemia show homozygous loss of atmgenes within the tumor cells.147 The malignant cells of approximately 35% of B-CLL patients harbour atm mutations and show decreased expression of the ATM protein. This subset of patients was characterized by a more aggressive form of disease compared to patients with normal ATM expression. In some cases, heterozygous mutations are present in all somatic cells, suggesting a genetic predisposition to the disease.148 149 The role of the ATM-encoded protein in regulating apoptosis via the p53 pathway (section 5) suggests that loss of its expression in some B-CLL may contribute to the resistance to apoptosis characteristic of this malignancy.

The Ph1 translocation [t(9;22)], is associated with CML and results in the fusion of the bcr and abl genes. The fusion gene encodes a 210-kD oncoprotein (p210bcr/abl) with enhanced protein tyrosine kinase activity compared with the normal abl-encoded protein.150 A variant Ph1 translocation associated with ALL encodes a 185-kD BCR/ABL oncoprotein (p185bcr/abl) with increased transforming potential compared with p210bcr/abl.151 

The normal abl-encoded protein is involved in the induction of apoptosis in some cell types.152 By contrast, expression of p210bcr/abl protects cells from killing by radiation, cytotoxic drugs, and ligation of FAS.153-155p210bcr/abl protects cells from killing by cytotoxic drugs by preventing the release of cytochrome c156 and activation of caspase 3.156 157 

Both p210bcr/abl and p185bcr/abl activate phosphatidylinositol 3′-kinase,158,159 and this pathway is essential for transformation by these oncoproteins.160 Kinase activation is mediated via binding of a complex containing the CRKL and CBL adaptor proteins to a proline-rich domain of both chimeric oncoproteins.150 161 Because activation of phosphatidylinositol 3-kinase results in phosphorylation of the pro-apoptotic BAD protein via protein kinase B (section 4.5), the anti-apoptotic actions of BCR/ABL oncoproteins may be mediated at least in part by this route.

Drugs that selectively inhibit the bcr/abl-encoded protein kinases may be of value in the treatment of CML and Ph1-positive ALL. Herbimycin A162 and CGP 57148163,164 selectively inhibit the expansion of cells and cell lines expressing bcr/abl oncoproteins. The actions of herbimycin A on Ph1-positive cell lines is markedly enhanced by combination with etoposide or γ radiation.165Antisense oligonucleotides that suppress expression of bcr/abloncoproteins may also enhance apoptosis induction by drugs or radiation.153 However, the actions of at least somebcr/abl antisense oligonucleotides on CML cell lines may be nonspecific.166 These nonspecific cytotoxic effects may be attributable to the release of deoxyribonucleotides as a consequence of exonucleolytic degradation of the oligonucleotides.167 

Chromosomal translocations associated with specific sub-types of acute leukemia and lymphoma result in the rearrangement of a variety of transcription factor genes. Some of these translocations result in enhanced expression of the transcription factor due to the juxtaposition of its gene next to Ig or T-cell antigen receptor loci, which contain powerful transcriptional enhancer elements. Other translocations involve the breakage and rejoining within introns of two transcription factor genes. The resulting hybrid genes encode chimeric transcription factors with novel properties which contribute to leukemogenesis.168 Some of these chimeric or aberrantly expressed proteins may contribute to malignant transformation via the suppression of apoptosis. Selected examples are described here.

11.1. Translocations resulting in inhibition of apoptosis.

The t(9;14) translocation, which is associated with 50% of lymphoplasmacytoid lymphoma, juxtaposes the paired box-containing genepax-5 to the IgH locus, where transcription from the pax-5promoters is augmented due to the proximity of the powerful Eμ enhancer.169,170 The PAX-5 protein represses transcription of the p53 gene.171 Therefore, decreased expression of p53 in cells bearing the t(9;14) translocation may contribute to malignant transformation via reducing expression of p53.

The chimeric PML-RARα transcription factor is generated as a consequence of juxtaposition of the retinoic acid receptor α gene (rarα; chromosome 17) and the pml gene (chromosome 15) in acute promyelocytic leukemia. Ectopic expression of this protein in myeloid cell lines diminishes apoptotic cell death. However, a reduced capacity to differentiate may also contribute to malignant transformation by the PML-RARα protein.172 

The t(17;19) translocation, which is associated with pre-B cell leukemia, results in the fusion of the genes encoding the transcription factors E2A and HLF (hepatic leukemia factor). Human leukemia cells expressing the chimeric E2A-HLF protein undergo rapid apoptotic death after ectopic expression of a dominant negative inhibitor of E2A-HLF function. In addition, ectopic expression of E2A-HLF in nonmalignant pro-B lymphocytes abrogates apoptosis induction induced by IL-3 withdrawal or by p53 expression.173 Therefore, the oncogenic action of E2A-HLF may be related to its ability to prevent apoptosis. The similarity of the DNA binding/dimerization domains of HLF to the CES-2 (cell death specification-2) protein of C elegans suggests that E2A-HLF may block apoptosis via inducing transcription of a gene whose protein product blocks an early step in the apoptotic pathway.173 

The t(10;14) translocation is associated with some cases of T-cell leukemia. The resulting juxtaposition of the hox 11 gene to the IgH locus results in its overexpression. Disruption of the hox 11 gene in mice results in the apoptotic death of spleen cells, again suggesting that oncogenic transformation by deregulated hox 11 expression is the result of protection from apoptosis.174 

Overexpression of the tal1(scl) gene as a result of the t(1;14) translocation is a frequent event in T-ALL. Ectopic expression of TAL1 in an immature human T-lymphoid cell line does not perturb cell-cycle control but results in a marked resistance to cytotoxic drugs and to FAS ligation.175 This anti-apoptotic action is dependent on the DNA-binding domain of TAL1, suggesting that induction of expression of an unknown gene(s) underlies the resistance of TAL1 overexpressing cells to cell killing.

11.2. Translocations resulting in induction of apoptosis.

By contrast, some transcription factors involved in chromosomal translocations induce apoptotic cell death. Expression of the MYC gene is deregulated as a result of juxtaposition to the IgH locus in Burkitt’s lymphoma and L3 ALL cells bearing the t(8;14) translocation.176 The ability of the MYC gene product to trigger cell-cycle transit contributes to malignant transformation induced by its overexpression. However, the MYC protein also induces apoptosis when ectopically expressed at high levels in fibroblasts. Although p53 is required for MYC-induced apoptosis,177 p53 “knockout” mice develop normally.178 Therefore, it is probable that levels of MYC protein generated during normal physiological responses do not induce apoptosis.

A net increase in cell number after enforced induction of MYC in murine fibroblasts requires that the apoptotic pathway be blocked by survival-inducing cytokines.179 Therefore, oncogenic transformation by deregulated MYC may require additional genetic events that abrogate apoptosis induction and may explain the high proportion of Burkitt’s lymphoma and L3 ALL cases bearing p53 gene lesions.137 The ability of overexpressed BCL-2 to collaborate with MYC in promoting the generation of lymphoid tumors in doubly transgenic mice is also consistent with the concept that oncogenenic transformation by MYC is dependent on the suppression of apoptosis.106 

The e2a-pbx1 fusion gene results from the t(1;19) translocation associated with B-cell precursor ALL. The protein product of this fusion gene rapidly induces apoptosis in B-cell progenitors. The dependence of apoptosis induction on the DNA-binding homeodomain of the PBX1 moiety suggests that cell killing is dependent on transcriptional activation of an unknown gene(s).180 Apoptosis induced by E2A-PBX1 expression in cell lines is blocked by BCL-2 expression. Therefore, oncogenic transformation by this chimeric gene may also depend on additional genetic changes that block apoptosis induction.180 

Knowledge of the complex biochemical pathways involved in the regulation of apoptosis in hematopoietic cells is advancing rapidly. Here we have focused on aspects of apoptosis regulation with particular relevance to the hematopoietic system. The BCL-2 protein family, the release of mitochondrial cytochrome c, p53-mediated transcriptional control, FAS, and the TNF receptor are involved in the control of apoptosis induction at least in some hematopoietic cells. Diverse regulatory mechanisms converge on a final common pathway involving activation of the caspase family of proteases. Additional interactions involving the BCL-2 family may also be important in apoptosis regulation and have been reviewed elsewhere.181 

Elevated expression of BCL-2, loss of functional p53, constitutive activativation of protein tyrosine kinases, or the generation of chimeric, oncogenic transcription factors as a result of chromosomal translocations abrogate apoptosis induction and antagonize the actions of cytotoxic drugs or of radiation on leukemia cells. Therefore, strategies designed to bypass blocks in the detection of apoptotic signals or in the signal transduction phase may be of value in overcoming resistance to cytotoxic regimes.2 However, apoptosis is a complex physiological process dependent on the integrated functioning of a large number of gene products. Therefore, any therapeutic stratagem that is dependent solely on the induction of apoptosis will lead to the rapid evolution of clones resistant to killing. Microenvironemental factors also influence the outcome of cytotoxic treatments by modulating pathways of apoptosis control. Therefore, manipulation of the cytokine levels of leukemia or lymphoma patients may also impact on the efficacy of chemotherapy or radiation.182 

Elucidation of the precise mechanisms involved in the apoptotic killing of leukemia cells (as opposed to cell lines) and of the strategies by which malignant cells escape killing by cytotoxic agents are major topics for future research. It is anticipated that an understanding of these facets of leukemia and lymphoma cell biology will lead to the design of effective strategies for the treatment of hematopoietic malignancies that are resistant to conventional treatment.

1
Fisher
DE
Apoptosis in cancer therapy: Crossing the threshold.
Cell
78
1994
539
2
Hannun
YA
Apoptosis and the dilemma of cancer chemotherapy.
Blood
89
1997
1845
3
Kerr
JFR
Winterford
CM
Harmon
BV
Apoptosis: Its significance in cancer and cancer therapy.
Cancer
73
1994
2013
4
Fadok
VA
Voelker
DR
Campbell
PA
Cohen
JJ
Bratton
DL
Henson
PM
Exposure of phosphatidylserine on the surface of apoptotic lymphocytes triggers specific recognition and removal by macrophages.
J Immunol
148
1992
2207
5
Martin
SJ
Reuterlingsperger
CP
McGahon
AJ
Rader
JA
van Schie
RC
LaFace
DM
Green
DR
Early redistribution of plasma membrane phosphatidylserine is a general feature of apoptosis regardless of the initiating stimulus: Inhibition by overexpression of bcl-2 and abl.
J Exp Med
182
1995
1545
6
Hengartner
MO
Horvitz
HR
Programmed cell death in Caenorhabditis elegans.
Curr Opin Genet Dev
4
1994
581
7
Hengartner
MO
Horvitz
HR
C. elegans cell-survival gene ced-9 encodes a functional homolog of the mammalian proto-oncogene bcl-2.
Cell
76
1994
665
8
Vaux
DL
Weissman
IL
Kim
SK
Prevention of programmed cell death in Caenorhabditis elegans by human bcl-2.
Science
258
1992
1955
9
Huang
DCS
Adams
JM
Cory
S
The conserved N-terminal BH4 domain of bcl-2 homologues is essential for inhibition of apoptosis and interaction with CED-4.
EMBO J
17
1998
1029
10
Thornberry
NA
Lazebnik
Y
Caspases: Enemies within.
Science
281
1998
12
11
Green
DR
Reed
JC
Mitochondria and apoptosis.
Science
281
1998
1309
12
Nicholson
DW
Ali
A
Thornberry
NA
Vaillancourt
JP
Ding
CK
Gallant
M
Gareau
Y
Griffin
PR
Labelle
M
Lazebnik
YA
Munday
NA
Raju
SM
Smulson
ME
Yamin
T-T
Yu
V
Miller
DK
Identification and inhibition of the ICE/CED-3 protease necessary for mammalian apoptosis.
Nature
376
1995
37
13
Tewari
M
Quan
LT
O’Rourke
K
Desnoyers
S
Zeng
Z
Beidler
DR
Poirier
GG
Salvesen
GS
Dixit
VM
Yama/CPP32β, a mammalian homolog of ced-3, is crmA-inhibitable protease that cleaves the death substrate poly(ADP-ribose) polymerase.
Cell
81
1995
801
14
Kothakota
S
Azuma
T
Reinhard
C
Klippel
A
Tang
J
Chu
K
McGarry
TJ
Kirschner
MW
Koths
K
Kwiatowski
DJ
Williams
LT
Caspase 3-generated fragment of gelsolin: Effector of morphological changes in apoptosis.
Science
278
1997
294
15
Liu
X
Zou
H
Slaughter
C
Wang
X
DFF, a heterodimeric protein that functions downstream of caspase 3 to trigger DNA fragmentation during apoptosis.
Cell
89
1997
175
16
Orth
K
Chinnaiyan
AM
Garg
M
Froelich
CJ
Dixit
VM
The ced-3/ICE-like protease Mch2 is activated during apoptosis and cleaves the death substrate lamin A.
J Biol Chem
271
1996
16443
17
Roy
N
Devereaux
QL
Takahashi
R
Salvesen
GS
Reed
JC
The cIAP-1 and c-IAP-2 proteins are direct inhibitors of specific caspases.
EMBO J
16
1997
6914
18
Vucic
D
Kaiser
WJ
Miller
LK
Inhibitor of apoptosis oroteins physically intract with and block apoptosis induced by Drosophila proteins HID and GRIM.
Mol Cell Biol
18
1998
3300
19
Ashkenazi
A
Dixit
VM
Death receptors: Signaling and modulation.
Science
281
1998
1305
20
Chinnaiyan
AM
O’Rourke
K
Tewari
M
Dixit
VM
FADD, a novel death domain-containing protein, interacts with the death domain of Fas and initiates apoptosis.
Cell
81
1995
505
21
Muzio
M
Chinnaiyan
AM
Kischkel
FC
O’Rourke
K
Shevchenko
A
Ni
J
Scaffidi
C
Bretz
JD
Zhang
M
Gentz
R
Mann
M
Krammer
PH
Peter
ME
Dixit
VM
FLICE, a novel FADD-homologous ICE/CED-3-like protease, is recruited to the CD95 (Fas/apo-1) death-inducing signaling complex.
Cell
85
1996
817
22
Boldin
MP
Goncharov
TM
Goltsev
YV
Wallach
D
Involvement of MACH, a novel MORT1/FADD-interacting protease, in Fas/Apo-1 and TNF receptor-induced cell death.
Cell
85
1996
803
23
Berke
G
The CTL’s kiss of death.
Cell
81
1995
9
24
Darmon
AJ
Nicholson
DW
Bleackley
RC
Activation of the apoptotic protease CPP32 by cyotoxic T cell-derived granzyme B.
Nature
377
1995
446
25
Martin
SJ
Amarantes-Mendes
GP
Shi
L
Chuang
TH
Casiano
CA
O’Brien
GA
Fitzgerald
P
Tan
EM
Bokoch
GM
Greenberg
AM
Green
DR
The cytotoxic cell protease granzyme B initiates apoptosis in a cell-free system by proteolytic processing and activation of the ICE/CED-3 family protease CPP32 via a novel two-step mechanism.
EMBO J
15
1996
2407
26
Froelich
CJ
Dixit
VM
Yang
X
Lymphocyte granule-mediated apoptosis: Matters of viral mimicry and deadly proteases.
Immunol Today
19
1998
30
27
Moretta
A
Molecular mechanisms in cell-mediated cytotoxicity.
Cell
90
1997
13
28
Vaux
DL
Cory
S
Adams
JM
Bcl-2 gene promotes haemopoietic cell survival and cooperates with c-myc to immortalize pre-B cells.
Nature
335
1988
440
29
Tsujimoto
Y
Stress-resistance conferred by high levels of bcl-2α protein in human B lymphoblastoid cell.
Oncogene
4
1989
1331
30
Yang
E
Korsmeyer
SJ
Molecular thanatopsis: A discourse on the bcl-2 family.
Blood
88
1996
386
31
Muchmore
SW
Sattler
M
Liang
H
Meadows
RP
Harlan
JE
Yoon
HS
Nettesheim
D
Chang
BS
Thompson
CB
Wong
S-L
Ng
S-C
Fesik
SW
X-ray and NMR structure of human Bcl-XL, an inhibitor of programmed cell death.
Nature
381
1996
335
32
Sattler
M
Liang
H
Nettesheim
D
Meadows
RP
Harlan
JE
Eberstadt
M
Yoon
HS
Shuker
SB
Chang
BS
Minn
AJ
Thompson
CB
Fesik
SW
Structure of bcl-XL-Bak peptide complex: Recognition between regulators of apoptosis.
Science
275
1997
983
33
Oltvai
ZN
Korsmeyer
SJ
Checkpoints of duelling dimers foil death wishes.
Cell
79
1994
189
34
Hsu
Y-T
Youle
RJ
Nonionic detergents induce dimerization among members of the bcl-2 famly.
J Biol Chem
272
1997
13829
35
Knudson
CM
Korsmeyer
SJ
Bcl-2 and Bax function independently to regulate cell death.
Nat Genet
16
1997
358
36
Simonian
PL
Grillot
DAM
Andrews
DW
Leber
B
Nunez
G
Bax homodimerization is not required for bax to accelerate chemotherapy-induced cell death.
J Biol Chem
271
1996
32073
37
Chen-Leavy
Z
Cleary
ML
Membrane topology of the Bcl-2 proto-oncogenic protein demonstrated in vitro.
J Biol Chem
265
1990
4929
38
Monaghan
P
Robertson
D
Amos
TA
Dyer
MJ
Mason
DY
Greaves
MF
Ultrastructural localization of bcl-2 protein.
J Histochem Cytochem
40
1992
1819
39
Wolter
KG
Hsu
Y-T
Smith
CL
Nechushtan
A
Xi
X-G
Youle
RJ
Movement of bax from cytosol to mitochondria during apoptosis.
J Cell Biol
139
1997
1281
40
Zou
H
Henzel
WJ
Liu
X
Lutschg
A
Wang
X
Apaf-1, a human protein homologous to C. elegans CED-4, participates in cytochrome c-dependent activation of caspase 3.
Cell
90
1997
405
41
Li
P
Nijhawan
D
Budihardjo
I
Srinivasula
SM
Ahmad
M
Alnemri
ES
Wang
X
Cytochrome c and dATP-dependent formation of Apaf-1/caspase-9 complex initiates an apoptotic protease cascade.
Cell
91
1997
479
42
Srinivasula
SM
Ahmad
M
Fernades-Alnemri
T
Alnemri
ES
Autoactivation of pro-caspase 9 by Apaf-1-mediated oligomerization.
Molec Cell
1
1998
949
43
Yang
J
Liu
X
Bhalla
K
Kim
CN
Ibrado
AM
Cai
J
Peng
T-I
Jones
DP
Wang
X
Prevention of apoptosis by bcl-2: Release of cytochrome c from mitochondria blocked.
Science
275
1997
1129
44
Kluck
RM
Bossy-Wetzel
E
Green
DR
Newmeyer
DD
The release of cytochrome c from mitochondria: A primary site for bcl-2 regulation of apoptosis.
Science
275
1997
1132
45
Vander Heiden
MG
Chandel
NS
Williamson
EK
Schumacker
PT
Thompson
CB
Bcl-XL regulates the membrane potential and volume homeostasis of mitochondria.
Cell
91
1997
627
46
Jurgensmeier
JM
Xie
Z
Devereaux
Q
Ellerby
L
Bredesden
D
Reed
JC
Bax directly induces release of cytochrome c from isolated mitochondria.
Proc Natl Acad Sci USA
95
1998
4997
47
Hengartner
MO
Death cycle and Swiss army knives.
Nature
391
1998
441
48
Minn
AJ
Velez
P
Schendel
SL
Liang
H
Muchmore
SW
Fesik
SW
Fill
M
Thompson
CB
Bcl-xL forms an ion channel in synthetic lipid membranes.
Nature
385
1997
353
49
Schendel
SL
Xie
Z
Montal
MO
Matsuyama
S
Montal
M
Reed
JC
Channel formation by anti-apoptotic protein bcl-2.
Proc Natl Acad Sci USA
94
1997
5113
50
Antonsson
B
Conti
F
Ciavatta
A
Montessuit
S
Lewis
S
Martinou
I
Bernasconi
L
Bernard
A
Mermod
J-J
Mazzei
G
Maundrell
K
Gambale
F
Sadoul
R
Martinou
J-C
Inhibition of bax channel-forming activity by bcl-2.
Science
277
1997
370
51
Marchetti
P
Castedo
M
Susin
SA
Zamzami
N
Hirsch
T
Macho
A
Haeffner
A
Hirsch
F
Geuskens
M
Kroemer
G
Mitochondrial permeability transition is a central coordinating event of apoptosis.
J Exp Med
184
1996
1155
52
Decaudin
D
Geley
S
Hirsch
T
Castedo
M
Marchetti
P
Macho
A
Kofler
R
Kroemer
G
Bcl-2 and bcl-XL antagonize the mitochondrial dysfunction preceding nuclear apoptosis induced by chemotherapeutic agents.
Cancer Res
57
1997
62
53
Shimizu
S
Eguchi
Y
Kamiike
W
Funahashi
Y
Mignon
A
Lacronique
V
Matsuda
H
Tsujimoto
Y
Bcl-2 prevents apoptotic mitochondrial dysfunction by regulating proton flux.
Proc Natl Acad Sci USA
95
1998
1455
54
Rosse
T
Olivier
R
Monney
L
Rager
M
Conus
S
Fellay
I
Jansen
B
Borner
C
Bcl-2 prolongs cell survival after bax-induced release of cytochrome c.
Nature
391
1998
496
55
Zhivotovsky
B
Orrenius
S
Brustugun
OT
Doskeland
SO
Injected cytochrome c induces apoptosis.
Nature
391
1998
449
56
Hu
Y
Benedict
MA
Wu
D
Inohara
N
Nunez
G
Bcl-XL interacts with Apaf-1 and inhibits Apaf-1-dependent caspase 9 activation.
Proc Natl Acad Sci USA
95
1998
4386
57
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
275
1997
1122
58
Marzo
I
Susin
SA
Petit
PX
Ravagnan
L
Brenner
C
Larochette
N
Zamzami
N
Kroemer
G
Caspases disrupt mitochondrial membrane barrier function.
FEBS Lett
427
1998
198
59
Zha
J
Harada
H
Yang
E
Jockel
J
Korsmeyer
SJ
Serine phosphorylation of death agonist BAD in response to survival factor results in binding to 14-3-3, not bcl-XL.
Cell
87
1996
619
60
del Peso
L
Gonzalez-Garcia
M
Page
C
Herrera
R
Nunez
G
Interleukin 3-induced phosphorylation of BAD through protein kinase Akt.
Science
278
1997
687
61
Datta
SR
Dudek
H
Tao
X
Masters
S
Fu
H
Gotoh
Y
Greenberg
ME
Akt phosphorylation of BAD couples survival signals to the cell-intrinsic death machinery
Cell
91
1997
231
62
Hemmings
BA
Akt signaling: Linking membrane events to life and death decisions.
Science
275
1997
628
63
Levine
AJ
p53, the cellular gatekeeper for growth and cell division.
Cell
88
1997
323
64
Kastan
MB
Zhan
Q
El-Deiry
WS
Carrier
F
Jacks
T
Walsh
WV
Plunkett
BS
Vogelstein
B
Fornace
AJ
A mammalian cell cycle checkpoint pathway utilizing p53 and GADD45 is defective in ataxia telangiectasia.
Cell
71
1992
587
65
Lee
S
Elenbaas
B
Levine
AJ
Griffith
J
p53 and its 14 kDa C-terminal domain recognize primary DNA damage in the form of insertion/deletion mismatches.
Cell
81
1995
1013
66
Lowe
SW
Schmitt
EM
Smith
SW
Osborne
BA
Jacks
T
p53 is required for radiation-induced apoptosis in mouse thymocytes.
Nature
362
1993
847
67
Clarke
AR
Purdie
CA
Harrison
DJ
Morris
RG
Bird
CC
Hooper
ML
Wyllie
AH
Thymocyte apoptosis induced by p53-dependent and independent pathways.
Nature
362
1993
849
68
Miyashita
T
Reed
JC
Tumor suppressor p53 is a direct transcriptional activator of the human bax gene.
Cell
80
1995
293
69
Miyashita
T
Krajewski
S
Krajewska
M
Wang
H-G
Lin
HK
Liebermann
DA
Hoffman
B
Reed
JC
Tumor suppressor p53 is a regulator of bcl-2 and bax gene expression in vitro and in vivo.
Oncogene
9
1994
1799
70
McCurrach
ME
Connor
TMF
Knudson
CM
Korsmeyer
SJ
Lowe
SW
Bax-deficiency promotes drug resistance and oncogenic transformation by attenuating p53-dependent apoptosis.
Proc Natl Acad Sci USA
94
1997
2345
71
Brady
HJM
Salomons
GS
Bobeldijk
RC
Berns
AJM
T cells from baxα transgenic mice show accelerated apoptosis in response to stimuli but do not show restored DNA damage-induced cell death in he absence of p53.
EMBO J
15
1996
1221
72
Polyak
K
Xia
Y
Zweier
JL
Kinzler
KW
Vogelstein
B
A model for p53-induced apoptosis.
Nature
389
1997
300
73
Zamzami
N
Susin
SA
Marchetti
P
Hirsch
T
Gomez-Monterrey
I
Castedo
M
Kroemer
G
Mitochondrial control of nuclear apoptosis.
J Exp Med
183
1996
1533
74
Johnson
TM
Yu
ZX
Ferrans
VJ
Lowenstein
RA
Finkel
T
Reactive oxygen species are downstream mediators of p53-dependent apoptosis.
Proc Natl Acad Sci USA
93
1996
11848
75
Jacobson
MD
Raff
MC
Programmed cell death and bcl-2 protection in very low oxygen.
Nature
374
1995
814
76
Muschel
RJ
Bernhard
EJ
Garza
L
McKenna
WG
Koch
CJ
Induction of apoptosis at different oxygen tensions: Evidence that oxygen radicals do not mediate apoptotic signaling.
Cancer Res
55
1995
995
77
Shen
Y
Shenk
T
Relief of p53-mediated transcriptional repression by the adenovirus E1B 19 kDa protein or the cellular bcl-2 protein.
Proc Natl Acad Sci USA
91
1994
8940
78
Caelles
C
Helmberg
A
Karin
M
p53-dependent apoptosis in the absence of transcriptional activation of p53 target genes.
Nature
370
1994
220
79
Dubrez
L
Goldwasser
F
Genne
P
Pommier
Y
Solary
E
The role of cell cycle regulation and apoptosis triggerring in determining the sensitivity of leukemic cells to topoisomerase I and II inhibitors.
Leukemia
9
1995
1013
80
Santana
P
Pena
LA
Haimovitz-Friedman
A
Martin
S
Green
D
McLaughlin
M
Schuchman
EH
Fuks
Z
Kolesnick
R
Acid sphingomyelinase-deficient human lymphoblasts are defective in radiation-induced apoptosis.
Cell
86
1996
189
81
Lowe
SW
Bodis
S
McClatchey
A
Remington
L
Ruley
HE
Fisher
DE
Housman
DE
Jacks
T
p53 status and the efficacy of cancer therapy in vivo.
Science
266
1994
807
82
Zhan
Q
Fan
S
Bae
I
Guillof
C
Liebermann
DA
O’ Connor
PM
Fornace
AJ
Induction of bax by genotoxic stress in human cells correlates with normal p53 status and apoptosis.
Oncogene
9
1995
3743
83
Jost
CA
Marin
MC
Kaelin
WG
p73 is a human p53-related protein that can induce apoptosis.
Nature
389
1997
191
84
Sachs
L
The control of hematopoiesis and leukemia: From basic biology to the clinic.
Proc Natl Acad Sci USA
93
1996
4742
85
Leary
AG
Zeng
HQ
Clark
SC
Ogawa
M
Growth factor requirements for survival in G0 and entry into the cell cycle of primitive human hemopoietic progenitors.
Proc Natl Acad Sci USA
89
1992
4013
86
Borge
OJ
Ramsfjell
V
Cui
L
Jacobsen
SE
Ability of early acting cytokines to directly promote survival and suppress apoptosis of human primitive CD34+CD38− bone marrow cells with multilineage potential at the single cell level: Key role of thrombopoietin.
Blood
90
1997
2282
87
Brandt
JE
Bhalla
K
Hoffman
R
Effects of interleukin-3 and c-kit ligand on the survival of various classes of human hematopoietic progenitor cells.
Blood
83
1994
1507
88
Muench
MO
Roncarolo
MG
Menon
S
Xu
Y
Kastelein
R
Zurawski
S
Hannum
CH
Culpepper
J
Lee
F
Namikawa
R
FLK-2/FLT-3 ligand regulates the growth of early myeloid progenitors isolated from human fetal liver.
Blood
85
1995
963
89
Veiby
OP
Jacobsen
FW
Cui
L
Lyman
SD
Jacobsen
SE
The flt3 ligand promotes the survival of primitive hemopoietic progenitor cells with myeloid as well as B lymphoid potential. Suppression of apoptosis and counteraction by TNF α and TGF-β.
J Immunol
157
1996
2953
90
Seleri
C
Maciejewski
JP
Sato
T
Young
NS
Interferon-γ constitutively expressed in the stromal microenvironement of human marrow cultures mediates potent hematopoietic inhibition.
Blood
87
1996
4149
91
Maciejewski
J
Selleri
C
Anderson
S
Young
NS
Fas antigen expression on CD34+ human marrow cells is induced by interferon γ and tumor necrosis factor α and potentiates cytokine-mediated hematopoietic suppression in vitro.
Blood
85
1995
3183
92
Manabe
A
Murti
KG
Coustan-Smith
E
Kumagai
M-A
Behm
FG
Raimondi
SC
Campana
D
Adhesion-dependent survival of normal and leukemic human B lymphoblasts on bone marrow stromal cells.
Blood
83
1994
758
93
Murti
KG
Brown
PS
Kumagai
M-A
Campana
D
Molecular interactions between B-cell progenitors and the bone marrow microenvironement.
Exp Cell Res
226
1996
47
94
Lotem
J
Sachs
L
Hematopoietic cells from mice deficient in wild-type p53 are more resistant to induction of apoptosis by some agents.
Blood
82
1993
1092
95
Blandino
G
Scardigli
R
Rizzo
MG
Crescenzi
M
Soddu
S
Sacchi
A
Wild-type p53 modulates apoptosis of normal, IL-3 deprived hematopoietic cells.
Oncogene
10
1995
731
96
Lotem
J
Sachs
L
Hematopoietic cytokines inhibit apoptosis induced by transforming growth factor β1 and cancer chemotherapy compounds in myeloid leukemic cells.
Blood
80
1992
1750
97
Lotem
J
Sachs
L
Interferon-γ inhibits apoptosis induced by wild-type p53, cytotoxic anticancer agents and viability factor deprivation in myeloid cells.
Leukemia
9
1995
685
98
Yonish-Rouach
E
Resnitzky
D
Lotem
J
Sachs
L
Kimchi
A
Oren
M
Wild-type p53 induces apoptosis of myelod leukaemic cells that is inhibited by interleukin-6.
Nature
352
1991
345
99
Nunez
G
Merino
R
Grillot
D
Gonzalez-Garcia
M
Bcl-2 and Bcl-x: Reglatory switches for lymphoid death and survival.
Immunol Today
15
1994
582
100
Dubrez
L
Savoy
I
Hamman
A
Solary
E
Pivotal role of a DEVD-sensitive step in etoposide-induced and fas-mediated apoptotic pathways.
EMBO J
15
1996
5504
101
Datta
R
Banach
D
Kojima
H
Talanian
RV
Alnemri
ES
Wong
WW
Kufe
DW
Activation of the CPP32 protease in apoptosis induced by 1-β-D arabinofuranosylcytosine and other DNA-damaging agents.
Blood
88
1996
1936
102
Belosillo
B
Dalmau
M
Colomer
D
Gil
J
Involvement of CED-3/ICE proteases in the apoptosis of B-chronic lymphocytic leukaemia cells.
Blood
89
1997
3378
103
Seto
M
Jaeger
U
Hockett
RD
Graninger
W
Bennett
S
Goldman
P
Korsmeyer
SJ
Alternative promoters and exons, somatic mutation and deregulation of the bcl-2-Ig fusion gene in lymphoma.
EMBO J
7
1988
123
104
Petrovic
AS
Young
RL
Hilgarth
B
Ambros
P
Korsmeyer
SJ
Jaeger
U
The Ig heavy chain 3′ end confers a posttranscriptional processing advantage to bcl-2-Igh fusion RNA in t(14;18) lymphoma.
Blood
91
1998
3952
105
McDonell
TJ
Deane
N
Platt
FM
Nunez
G
Jaeger
U
McKearn
JP
Korsmeyer
SJ
bcl-2-immunoglobulin transgenic mice demonstrate extended B cell survival and follicular lymphoproliferation.
Cell
57
1989
79
106
Strasser
A
Harris
AW
Bath
ML
Cory
S
Novel primitive lymphoid tumours induced in transgenic mice by cooperation between myc and bcl-2.
Nature
348
1990
331
107
Miyashita
T
Reed
JC
Bcl-2 oncoprotein blocks chemotherapy-induced apoptosis in a human leukemia cell line.
Blood
81
1993
151
108
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
52
1992
5407
109
Minn
AJ
Rudin
CM
Boise
LH
Thompson
CB
Expression of bcl-xL can confer a multidrug resistant phenotype.
Blood
86
1995
1903
110
Rohatiner
A
Lister
TA
Management of follicular lymphoma.
Curr Opin Oncol
6
1994
473
111
Ghia
P
Boussiotis
VA
Schultze
JL
Cardodo
AA
Dorfman
DM
Gribben
JG
Freedman
AS
Nadler
LM
Unbalanced expression of bcl-2 family proteins in follicular lymphoma: Contribution of CD40 signalling in promoting survival.
Blood
91
1998
244
112
Webb
A
Cunningham
D
Cotter
F
Clarke
PA
di Stefano
F
Ross
P
Corbo
M
Dziewanowska
Z
Bcl-2 antisense therapy in patients with non-Hodgkin’s lymphoma.
Lancet
349
1997
1137
113
Plumas
J
Jacob
M-C
Chaperot
L
Molens
JP
Sotto
J-J
Bensa
J-C
Tumor B cells from non-Hodgkin’s lymphoma are resistant to CD95 (Fas/apo-1) mediated apoptosis.
Blood
91
1998
2875
114
Foon
KA
Rai
KR
Gale
RP
Chronic lymphocytic leukemia: New insights into biology and therapy.
Ann Intern Med
113
1990
525
115
Dyer
MJS
Zani
VJ
Lu
WZ
O’Byrne
A
Mould
S
Chapman
R
Heward
JM
Kayano
H
Jadayel
D
Matutes
E
Catovsky
D
Oscier
DG
Bcl-2 translocations in leukemias of mature B cells.
Blood
83
1994
3682
116
Thomas
A
El Rouby
S
Reed
JC
Krajewski
S
Silber
R
Potmesil
M
Newcomb
EW
Drug-induced apoptosis in B-cell chronic lymphocytic leukemia: Relationship between p53 gene mutations and bcl-2/bax proteins in drug resistance.
Oncogene
12
1996
1055
117
Pepper
C
Hoy
T
Bentley
DP
Bcl-2/bax ratios in chronic lymphocytic leukaemia and their correlation with in vitro and clinical drug resistance.
Br J Cancer
76
1997
935
118
Panayiotidis
P
Ganeshaguru
K
Jabbar
SAB
Hoffbrand
AV
Interleukin-4 inhibits apoptotic cell death and loss of bcl-2 protein in B-chronic lymphocytic leukaemia cells in vitro.
Br J Haematol
85
1993
439
119
Dancescu
M
Rubio-Trujillo
M
Biron
G
Bron
D
Delespesse
G
Sarfati
M
Interleukin-4 protects chronic lymphocytic leukemic B cells from death by apoptosis and upregulates bcl-2 expression.
J Exp Med
176
1992
1319
120
Panayiotidis
P
Ganeshaguru
K
Jabbar
SAB
Hoffbrand
AV
Alpha interferon (α-IFN) protects B-chronic lymphocytic leukaemia cells from apoptotic death in vitro.
Br J Haematol
86
1994
169
121
Jewell
AP
Lydyard
PM
Worman
CA
Giles
FJ
Goldstone
AH
Growth factors can protect B-chronic lymphocytic leukaemia cells against programmed cell death without promoting proliferation.
Leuk Lymphoma
18
1995
159
122
Buschle
M
Campana
D
Carding
SR
Richard
C
Hoffbrand
AV
Brenner
MK
Interferon-γ inhibits apoptotic cell death in B chronic lymphocytic leukemia.
J Exp Med
177
1993
213
123
Panayiotidis
P
Jones
D
Ganeshaguru
K
Foroni
L
Hoffbrand
AV
Human bone marrow stromal cells prevent apoptosis and support the survival of chronic lymphocytic leukaemia cells in vitro.
Br J Haematol
92
1996
97
124
Lagneaux
L
Delforge
A
Bron
D
De Bruyn
C
Stryckman
P
Chronic lymphocytic leukemic B cells but not normal B cells are rescued from apoptosis by contact with normal bone marrow stromal cells.
Blood
91
1998
2387
125
Kitada
S
Andersen
J
Akar
S
Zapata
JM
Takayama
S
Krajewski
S
Wang
H-G
Zhang
X
Bullrich
F
Croce
CM
Rai
K
Hines
J
Reed
JC
Expression of apoptosis-regulating proteins in chronic lymphocytic leukemia: Correlations with in vitro and in vivo chemoresponses.
Blood
91
1998
3379
126
Campos
L
Rouault
JP
Sabido
O
Oriol
P
Roubi
N
Vasselon
C
Archimbaud
E
Magaud
J-P
Guyotat
D
High expression of bcl-2 protein in acute myeloid leukemia cells is associated with poor response to chemotherapy.
Blood
81
1993
3091
127
Maung
ZT
MacLean
FR
Reid
MM
Pearson
ADJ
Proctor
SJ
Hamilton
PJ
Hall
AG
The relAtionship between bcl-2 expression and response to chemotherapy in acute leukaemia.
Br J Haematol
88
1994
105
128
Banker
DE
Groudine
M
Norwood
T
Appelbaum
FR
Measurement of spontaneous and therapeutic agent-induced apoptosis with bcl-2 protein expression in acute myeloid leukemia.
Blood
89
1997
243
129
Kaufmann
SH
Karp
JE
Svingen
PA
Krajewski
S
Burke
PJ
Gore
SD
Reed
JC
Elevated expression of the apoptotic regulator Mcl-1 at the time of leukemic relapse.
Blood
91
1998
991
130
Keith
FJ
Bradbury
DA
Zhu
Y-M
Russell
NH
Inhibition of bcl-2 with antisense oligonucleotides induces apoptosis and increases the sensitivity of AML blasts to Ara-C.
Leukemia
9
1995
131
131
Campos
L
Sabido
O
Sebban
C
Charrin
C
Bertheas
MF
Fiere
D
Guyotat
D
Expression of the bcl-2 proto-oncogene in adult acute lymphoblastic leukemia.
Leukemia
10
1996
434
132
Prokocimer
M
Rotter
V
Structure and function of p53 in normal cells and their aberrations in cancer cells: Projections on the hematologic cell lineages.
Blood
84
1994
2391
133
Ahuja
H
Bar-Eli
M
Arlin
Z
Advani
S
Allen
SL
Goldman
J
Snyder
D
Foti
A
Cline
M
The spectrum of molecular alterations in the evolution of chronic myelocytic leukemias.
J Clin Invest
87
1991
2042
134
El Rouby
S
Thomas
A
Costin
D
Rosenberg
CR
Potmesil
M
Silber
R
Newcomb
EW
p53 gene mutation in B-cell chronic lymphocytic leukemia is associated with drug resistance and is independent of MDR1/MDR3 gene expression.
Blood
82
1993
3452
135
Wattel
E
Preudhomme
C
Hecquet
B
Vanrumbeke
M
Quesnel
B
Dervitte
I
Morel
P
Fenaux
P
p53 mutations are associated with resistance to chemotherapy and short survival in hematologic malignancies.
Blood
84
1994
3148
136
Cordone
I
Masi
S
Mauro
FR
Soddu
S
Morsilli
O
Valentini
T
Vegna
ML
Guglielmi
C
Mancini
F
Giuliacci
S
Sacchi
A
Mandelli
F
Foa
R
p53 expression on B-cell chronic lymphocytic leukemia: A marker of disease progression and poor prognosis.
Blood
91
1998
4342
137
Gaidano
G
Ballerini
P
Gong
JZ
Inghirami
G
Neri
A
Newcomb
EW
Magrath
IT
Knowles
DM
Dalla-Favera
R
p53 mutations in human lymphoid malignancies: Association with Burkitt lymphoma and chronic lymphocytic leukemia.
Proc Natl Acad Sci USA
88
1991
5413
138
Sander
CA
Yano
T
Clark
HM
Harris
C
Longo
DL
Jaffe
ES
Raffeld
M
p53 mutation is associated with progression in follicular lymphomas.
Blood
82
1993
1994
139
Nakai
H
Tanaka
S
Nishigaki
H
Taniwaki
M
Yokota
S
Horiike
S
Takashima
T
Seriu
T
Nakagawa
H
p53 gene mutation and loss of a chromosome 17p in Philadelphia chromosome (Ph1)-positive acute leukemia.
Leukemia
7
1993
1547
140
Wada
M
Bartram
CR
Nakamura
H
Hachiya
M
Chan
D-L
Borenstein
J
Miller
CW
Ludwig
L
Hansen-Haggo
TE
Ludi
WD
Reiter
A
Mizoguchi
H
Koeffler
HP
Analysis of p53 mutations in a large series of lymphoid hematologic malignancies of childhood.
Blood
82
1993
3163
141
Fenaux
P
Jonveaux
P
Quiquandon
I
Lai
JL
Pignon
JM
Loncheux-Lefebvre
MH
Bauters
F
Berger
R
Kerckaert
JP
p53 gene mutations in acute myeloid leukemia with 17p monosomy.
Blood
78
1991
1652
142
Griffin
JD
Lowenberg
B
Clonogenic cells in acute myeloid leukemia.
Blood
68
1986
1185
143
Zhu
Y-M
Bradbury
DA
Russell
NH
Wild-ype p53 is required for apoptosis induced by growth factor deprivation in factor-dependent leukaemic cells.
Br J Cancer
69
1994
468
144
Kaplinsky
C
Lotem
J
Sachs
L
Protection of human myeloid leukemic cells against doxorubicin-induced apoptosis by granuloctye-macrophage colony-stimulating factor and interleukin 3.
Leukemia
10
1996
460
145
Bischoff
JR
Kirn
DH
Williams
A
Heise
C
Horn
S
Muna
M
Ng
L
Nye
JA
Sampson-Johannes
A
Fattaey
A
McCormick
F
An adenovirus mutant that replicates selectively in p53-deficient human cells.
Science
274
1996
342
146
Taylor
AM
Metcalfe
JA
Thick
J
Mak
YF
Leukemia and lymphoma in ataxia telangiectasia.
Blood
87
1996
423
147
Vorechovsky
I
Luo
L
Dyer
MJS
Catovsky
D
Amlot
PL
Yaxley
JC
Foroni
L
Hammarstrom
L
Webster
ADB
Yuille
MAR
clustering of missense mutations in the ataxia-telangiectasia gene in a sporadic T-cell leukemia.
Nature Genet
17
1997
96
148
Starostik
P
Manshouri
T
O’Brien
S
Freireich
E
Kantarjian
H
Haidar
M
Lerner
S
Keating
M
Albitar
M
Deficiency of the ATM protein expression defines an aggressive subgroup of B-cell chronic lymphocytic leukemia.
Cancer Res
58
1998
4552
149
Stankovic
T
Weber
P
Stewart
G
Bedenham
T
Murray
J
Byrd
PJ
Moss
PAH
Taylor
AMR
Inactivation of ataxia telangiectasia mutated gene in B-cell chronic lymphocytic leukemia.
Lancet
353
1999
26
150
Riordan FA, Wickremasinghe RG: Signal transduction by the Philadelphia chromosome-encoded BCR/ABL oncoproteins: Therapeutic implications for chronic myeloid leukemia and Philadelphia-positive acute lymphoblastic leukemia. Hematology 1999 (in press)
151
Kurzrock
R
Gutterman
JU
Talpaz
M
The molecular genetics of Philadelphia chromosome-positive leukemia.
N Engl J Med
319
1988
990
152
Yuan
Z-M
Huang
Y
Ishiko
T
Kharbanda
S
Weichselbaum
RR
Kufe
D
Regulation of DNA-damage-induced apoptosis by the c-Abl tyrosine kinase.
Proc Natl Acad Sci USA
94
1997
1437
153
McGahon
A
Bissonnette
R
Schmitt
M
Cotter
KM
Green
DR
Cotter
TM
Bcr-abl maintains resistance of chronic myelogenous leukemia cells to apoptotic cell death.
Blood
83
1994
1179
154
McGahon
AJ
Nishioka
WK
Martin
SJ
Mahboubi
A
Cotter
TG
Green
DR
Regulation of the fas apoptotic cell death pathway by abl.
J Biol Chem
270
1995
2625
155
Bedi
A
Zehnbauer
BA
Barber
JP
Sharkis
SJ
Jones
RJ
Inhibition of apoptosis by bcr-abl in chronic myeloid leukemia.
Blood
83
1994
2038
156
Amarantes-Mendes
G
Kim
CN
Liu
L
Huang
Y
Perkins
CL
Green
DR
Bhalla
K
Bcr-abl exerts its effect against diverse apoptotic stimuli through blockage of mitochondrial release of cytochrome c and activation of caspase 3.
Blood
91
1998
1700
157
Dubrez
L
Eymin
B
Sordet
O
Droin
N
Turhan
AG
Solary
E
BCR-ABL delays apoptosis upstream of procaspase 3 activation.
Blood
91
1998
2415
158
Griffin
JD
Carlesso
N
Matulonis
U
Ernst
T
Matsuguchi
T
Druker
B
The p210bcr/abl oncogene activates IL-3/GM-CSF receptor signal transduction pathways in myeloid cell lines.
Exp Hematol
21
1993
1010
159
Gotoh
A
Miyazawa
K
Ohyashiki
K
Toyama
K
Potential molecules implicated in downstream signaling pathways of p185bcr-abl in Ph+ ALL involves GTPase activating protein, phospholipase C-γ1 and phosphatidylinositol 3′-kinase.
Leukemia
8
1994
115
160
Skorski
T
Bellacosa
A
Nieborowska-Skorska
M
Majewski
M
Martinez
R
Choi
JK
Trotta
R
Wlodarski
P
Perrotti
D
Chan
TO
Wasik
MA
Tsichlis
PN
Calabretta
B
Transformation of hematopoietec cells by BCR/ABL requires activation of a PI-3K/Akt dependent pathway.
EMBO J
16
1997
6151
161
Sattler
M
Salgia
R
Okuda
K
Uemura
N
Durstin
MA
Pisick
E
Li
JL
Prasad
KV
Griffin
JD
The proto-oncogene product p210CBL and the adaptor proteins CRKL and c-CRK link c-ABL, p190BCR/ABL and p210BCR/ABL to the phosphatidylinositol 3′-kinase pathway.
Oncogene
12
1996
839
162
Okabe
M
Uehara
Y
Miyagishima
T
Itaya
T
Tanaka
M
Kuni-Eda
Y
Kurosawa
M
Miyazaki
T
Effect of herbimycin A, an antagonist of tyrosine kinase, on bcr/abl oncoprotein-associated cell proliferations.
Blood
80
1992
1330
163
Carroll
M
Ohno-Jones
S
Tamura
S
Buchdunger
E
Zimmermann
J
Lydon
NB
Gilliland
DG
Druker
BJ
CGP 57148, a tyrosine kinase inhibitor, inhibits the growth of cells expressing BCR-ABL, TEL-ABL and TEL-PDGFR fusion proteins
Blood
90
1997
4947
164
Deininger
MW
Goldman
JM
Lydon
N
Melo
JV
The tyrosine kinase inhibitor CGP57148B selectively inhibits the growth of BCR-ABL-positive cells.
Blood
90
1997
3691
165
Riordan
FA
Bravery
CA
Mengubas
K
Ray
N
Borthwick
NJ
Akbar
AN
Hart
SM
Hoffbrand
AV
Mehta
AB
Wickremasinghe
RG
Herbimycin A accelerates the induction of apoptosis following etoposide treatment or γ-irradiation of bcr/abl-positive leukaemia cells.
Oncogene
16
1998
1533
166
O’Brien
SG
Kirkland
MA
Melo
JV
Rao
MH
Davidson
RJ
McDonald
C
Goldman
JM
Antisense BCR-ABL oligomers cause non-specific inhibition of chronic myeloid leukemia cell lines.
Leukemia
8
1994
2156
167
Vaerman
JL
Moureau
P
Deldime
F
Lewalle
P
Lammineur
C
Morschhauser
F
Martiat
P
Antisense oligodeoxyribonucleotides suppress hematologic cell growth through stepwise release of deoxyribonucleotides.
Blood
90
1997
331
168
Look
AT
Oncogenic transcription factors in the human acute leukemias.
Science
278
1997
1059
169
Busslinger
M
Klix
N
Pfeffer
P
Graninger
PG
Kozmik
Z
Deregulation of PAX-5 by translocation of the Eμ enhancer of the IgH locs adjacent to two alternative PAX-5 promoters in a diffuse large-cell lymphoma.
Proc Natl Acad Sci USA
93
1996
6129
170
Iida
S
Rao
PH
Nallasivam
P
Hibshoosh
H
Butler
M
Louie
DC
Dyomin
V
Chaganti
RS
Dalla-Favera
R
The t(9;14)(p13;q32) chromosomal translocation associated with lymphoplasmacytoid lymphoma involves the PAX-5 gene.
Blood
88
1996
4110
171
Stuart
ET
Haffner
R
Oren
M
Gruss
P
(1995) Loss of p53 function through PAX-mediated transcriptional repression.
EMBO J
14
1995
5638
172
Grignani
F
Ferrucci
PF
Testa
U
Talamo
G
Fagioli
M
Alcalay
M
Mencarelli
A
Grignani
F
Peschle
C
Nicoletti
I
The acute promyelocytic leukemia-specific PML-RARα fusion protein inhibits differentiation and promotes survival of myeloid precursor cells.
Cell
74
1993
423
173
Inaba
T
Inukai
T
Yoshihara
T
Seyschab
H
Ashmun
RA
Canman
CE
Laken
SJ
Kastan
MB
Look
AT
Reversal of apoptosis by the leukaemia-associated E2A-HLF chimaeric transcription factor.
Nature
382
1996
541
174
Dear
TN
Colledge
WH
Carlton
MB
Lavenir
I
Larson
T
Smith
AJ
Warren
AJ
Evans
MJ
Sofroniew
MV
Rabbitts
TH
The Hox 11 gene is essential for cell survival during splenic development.
Development
121
1995
2909
175
Bernard
M
Delabesse
E
Novault
S
Hermine
O
MacIntyre
EA
Anti-apoptotic effect of ectopic TAL1/SCL expression in a human leukemic T cell line.
Cancer Res
58
1998
2680
176
Rabbitts
TH
Chromosomal translocations in human cancer.
Nature
372
1994
143
177
Hermeking
H
Eick
D
Mediation of c-myc-induced apoptosis by p53.
Science
265
1994
2091
178
Donehower
LA
Harvey
M
Slagle
BL
McArthur
MJ
Montgomery
CA
Butel
JS
Bradley
A
Mice deficient for p53 are developmentally normal but are susceptible to spontaneous tumors.
Nature
365
1992
215
179
Harrington
EA
Bennett
MR
Fanidi
A
Evan
GI
c-myc-induced apoptosis in fibroblasts is inhibited by specific cytokines.
EMBO J
13
1994
3286
180
Smith
KS
Jacobs
Y
Chang
CP
Cleary
ML
Chimeric oncoprotein E2a-Pbx 1 induces apoptosis of hematopoietic cells by a p53-independent mechanism that is suppressed by bcl-2.
Oncogene
14
1997
2917
181
Reed
JC
Double identity for proteins of the bcl-2 family.
Nature
387
1997
773
182
Sachs
L
Lotem
J
Control of programmed cell death in normal and leukemic cells: New implications for therapy.
Blood
82
1993
15
183
Ashkenazi
A
Dixit
VM
Death receptors: Signaling and modulation.
Science
281
1998
1305
184
Banin
S
Moyal
L
Shieh
S-Y
Taya
Y
Anderson
CW
Chessa
L
Smorodinsky
NI
Prives
C
Reiss
Y
Shiloh
Y
Ziv
Y
Enhanced phosphorylation of p53 by ATM in response to DNA damage.
Science
281
1998
1674
185
Canman
CE
Lim
D-S
Cimprich
KA
Taya
Y
Tamai
K
Sakaguchi
K
Appella
E
Kastan
MB
Siliciano
JD
Activation of the ATM kinase by ionizing radiation and phosphorylation of p53.
Science
281
1998
1677
186
Cardone
MH
Roy
N
Stennicke
HR
Salvesen
GS
Franke
TF
Stanbridge
E
Frisch
S
Reed
JC
Regulation of cell death protease caspase-9 by phosphorylation.
Science
282
1998
1318
187
Brunet
A
Bonni
A
Zigmond
MJ
Lin
MZ
Juo
P
Hu
LS
Anderson
MJ
Arden
KC
Blenis
J
Greenberg
ME
Akt promotes cell survival by phosphorylating and inhibiting a forkhead transcription factor.
Cell
96
1999
857
188
Willis
TG
Jadayel
DM
Du
M-Q
Peng
H
Perry
AR
Abdul-Rauf
M
Price
H
Karran
L
Majekodumni
O
Wlodarska
I
Pan
L
Crook
T
Hamoudi
R
Isaacson
DG
Dyer
MJS
Bcl10 is involved in t(1:14)(p22;q23) of MALT B cell lymphoma and mutated in multiple tumor types.
Cell
96
1999
35
189
Earnshaw
WC
A cellular poison cupboard.
Nature
397
1999
387

(1) New members of the cell-surface death receptor (DR) family have been described. DR3 triggers apoptosis consequent to binding APO 3 ligand.183 TNF-related apoptosis-inducing ligand (TRAIL) initiates cell death following binding to DR4 or DR5. The actions of DR4 and 5 are limited by decoy receptors which bind TRAIL but are unable to transduce apoptotic signals.183 (2) Binding of the MDM2 protein to p53 targets the latter for degradation. Phosphorylation of p53 by the ATM protein kinase results in its dissociation from MDM2 and consequent stabilization.184,185(3) Phosphorylation and inactivation of caspase 9186 and of the forkhead family transcription factor FKHRL1187 by protein kinase B contribute to the suppression of apoptosis by pro-survival cytokines. (4) The bcl10 gene involved in the t(1;14)(p22;q23) translocation of mucosa-associated lymphoid tissue (MALT) lymphoma has been cloned. Translocation results in expression of a truncated protein which lacks the pro-apoptotic activity of the wild-type BCL10 protein.188 (5) Mitochondria of cells primed for apoptosis can release death-inducing proteins other than cytochrome c. These include caspases 2 and 9 and a flavoprotein, apoptosis-inducing factor (AIF).189 

Author notes

Address reprint requests to R. Gitendra Wickremasinghe, PhD, Department of Hematology, Royal Free and University College Medical School, Rowland Hill St, London NW3 2PF, UK.

Sign in via your Institution