Our previous work showed that the nuclear scaffold (NS) protease is required for apoptosis of both thymocytes and chronic lymphocytic leukemic (CLL) lymphocytes. Because partial sequencing of one of the subunits of the NS protease revealed homology to the proteasome, we tested the effects of classical proteasome inhibitors on apoptosis in CLL cells. Here we report that proteasome inhibition caused high levels of DNA fragmentation in all patients analyzed, including those resistant to glucocorticoids or nucleoside analogs, in vitro. Proteasome inhibitor-induced DNA fragmentation was associated with activation of caspase/ICE family cysteine protease(s) and was blocked by the caspase antagonist, zVADfmk. Analysis of the biochemical mechanisms involved showed that proteasome inhibition resulted in mitochondrial dysregulation leading to the release of cytochrome c and a drop in mitochondrial transmembrane potential (▵Ψ). These changes were associated with inhibition of NFκB, a proteasome-regulated transcription factor that has been implicated in the suppression of apoptosis in other systems. Together, our results suggest that drugs that target the proteasome might be capable of bypassing resistance to conventional chemotherapy in CLL.

CHRONIC LYMPHOCYTIC leukemia (CLL) is an illness characterized by an accumulation of monoclonal mature B cells in the peripheral blood. Although CLL is the most common leukemia in the Western world, little is known about the biology of the disease. Treatment schemes rely heavily on glucocorticoids, chlorambucil, and nucleoside analogs, and we and others have shown that all of these agents trigger apoptosis in CLL cells in vitro, suggesting that induction of apoptosis may account for their therapeutic efficacy. Furthermore, recent work has shown that apoptosis in vitro correlates with Rai stage,1,2 and rates of apoptosis detected following fludarabine treatment correlate with clinical response in vivo.3 However, despite the initial effectiveness of these drugs in patients with low-grade disease, resistant cells ultimately emerge, leaving no effective treatment options available. It is possible that drug-resistant CLL cells possess intrinsic defect(s) in their ability to undergo apoptosis.

Protease activation is required for completion of the apoptotic program in all cellular and cell-free systems interrogated to date.4,5 Of central importance are members of the ICE/caspase family of aspartate-specific cysteine proteases, which appear to function at the core of the “effector” machinery for cell death. Caspase activation can either be directly promoted by oligomerization of certain caspase-associated cell surface “death” receptors (Fas, TNF-RI)6 or by intramitochondrial events that lead to the release of the electron transport chain intermediate, cytochrome c.7 Precisely how caspases promote the downstream features of apoptosis is not clear, but studies with specific peptide-based active site inhibitors indicate that they are required for all of the major biochemical events observed in apoptotic cells, including changes in cellular morphology, loss of plasma membrane asymmetry (exposure of phosphatidylserine on the outer leaflet), and DNA fragmentation.8 

We and others have obtained evidence that certain noncaspase proteases are also required for DNA fragmentation and apoptosis. Specifically, we have shown that peptide-based active site inhibitors of a Ca2+-dependent nuclear protease, termed the nuclear scaffold (NS) protease, block glucocorticoid- and nucleoside analog–induced DNA fragmentation in CLL lymphocytes.9Although the molecular characteristics of the NS protease are at present unclear, preliminary evidence obtained by another laboratory10 suggests that it is structurally and functionally related to the 26S multicatalytic protease complex (MPC), otherwise known as the proteasome. This possible similarity may explain why NS protease inhibitors block DNA fragmentation, because previous studies have implicated the proteasome in the programmed cell death of intersegmental muscles in the moth, Manduca sexta,11 and more recent work in isolated mouse thymocytes12 and neuronal cells13 has shown that proteasome inhibitors block caspase activation and other downstream events associated with apoptosis in these cells.

The results presented above suggested to us that the effects of NS protease inhibitors in CLL cells might be due to proteasome inhibition. To directly address this possibility, we tested the effects of several specific proteasome inhibitors on caspase activation and DNA fragmentation in isolated CLL lymphocytes, expecting that they would suppress apoptotic cell death. On the contrary, here we report that proteasome inhibition resulted in extraordinarily high levels of DNA fragmentation in all patient isolates analyzed, including those found to be completely resistant to glucocorticoid-induced apoptosis. Analysis of the biochemical mechanisms involved showed that the effects are linked to inhibition of NFκB, a transcription factor implicated in the maintenance of cell survival in other model systems.14-17 

Materials.

The esterified peptide caspase inhibitor, Z-VAD (OMe)fmk, the fluorigenic caspase substrate, DEVD-AMC, and the mouse anti-PARP monoclonal antibody C2-10 were purchased from Enzyme Systems Products, Inc (Dublin, CA). A peptide inhibitor of the NS protease, Z-APFcmk, and the caspase antagonist, Boc-Asp-chloromethylketone (BDcmk) were purchased from Bachem Bioscience (King of Prussia, PA). Monoclonal antibodies for caspase-3, p53, p27, and c-Jun were purchased from Transduction Laboratories (Lexington, KY). A monoclonal antibody to c-Fos and the proteasome inhibitors lactacystin and MG-132 were obtained from Calbiochem (San Diego, CA). Horseradish peroxidase-conjugated anti-mouse and anti-rabbit antibodies were from Amersham Corp (Arlington Heights, IL).

Patients, cell isolation, and incubation criteria.

All patients fulfilled the National Cancer Institute’s (NCI) criteria for the diagnosis of CLL. Some of the patients had received prior therapy, although none within the last 6 months before experimentation. Immunophenotyping by dual-parameter flow cytometry showed coexpression of CD5 with B-cell antigen and isotypic light chain expression. Clinical staging was based on the system described by Rai.18 Freshly isolated peripheral blood was fractionated by Ficoll-Hypaque (Winthrop Pharmaceuticals, New York, NY) sedimentation at 4°C. Nonadherent mononuclear cells were then immediately suspended in complete RPMI 1640 medium supplemented with 10% fetal calf serum (FCS), 10 mmol/L HEPES (pH 7.5), and antibiotics at a cellular concentration of 1 to 2 × 106 cells/mL. Cell viability was assessed by Trypan blue exclusion and exceeded 95% following the isolation procedure.

Granulocyte-colony stimulating factor (G-CSF)–mobilized progenitor cells were obtained from pheresis samples by magnetic cell sorting (MACS). The pheresis samples were resuspended in 50 mL of cold RPMI medium, and two “soft-spins” (200g, 10 minutes) were performed to remove platelets. Cells were labeled with anti-CD34 antibody and isolated with a commercial CD34 isolation kit (Miltenyi Biotec, Auburn, CA) according to the manufacturer’s instructions. A MACS buffer consisting of Ca2+/Mg2+-free Hanks’ buffered salt solution (HBSS) containing 0.6% ACD-A (Baxter, Deerfield, IL), 0.5% bovine serum albumin (BSA; Sigma, St Louis, MO), pH 6.5, was used throughout staining and separation to prevent cell clumping while maintaining optimal progenitor viability. Cells were separated on VS-positive selection columns using a VarioMACS according to the manufacturer’s instructions. Cell purity was assessed by flow cytometry using CD34-phycoerythrin (PE) and CD45-fluorescein isothiocyanate (FITC) as described previously.19 

DNA fragmentation analysis.

Quantification of apoptosis by propidium iodide (PI) staining and fluorescence-activated cell sorting (FACS) analysis was performed as described previously.20 Following incubation with various agents in vitro, cells were pelleted by centrifugation and resuspended in phosphate-buffered saline (PBS) containing 50 μg/mL PI, 0.1% Triton X-100, and 0.1% sodium citrate. Samples were stored at 4°C for 16 hours and vortexed before FACS analysis (FL-3 channel).

Cytochrome c release measurements.

Release of cytochrome c from mitochondria was measured by immunoblotting essentially as described previously.21 Cells were incubated in the absence or presence of 10 μmol/L APFcmk, 10 μmol/L MG-132, or 10 μmol/L methylprednisolone for 4 hours, obtained by centrifugation, and gently lysed for 30 seconds in an ice-cold buffer containing 250 mmol/L sucrose, 1 mmol/L EDTA, 0.1% digitonin, and 25 mmol/L Tris, pH 6.8. Lysates were centrifuged for 2 minutes at 12,000g, supernatants were mixed with 2× Laemmli’s reducing sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) sample buffer, and extracts from equal numbers of cells (10 to 20 × 106) were resolved by 15% SDS-PAGE. Polypeptides were transferred to nitrocellulose membranes (0.2 μm; Schleicher & Scheull, Keene, NH), and cytochrome c was detected by immunoblotting with a monoclonal antibody (clone 7H8.2C12; purchased from Pharmingen, San Diego, CA).

Caspase activity assay.

Protease activity measurements were conducted as described previously.9 Cells were lysed in 1 mL of a buffer containing 25 mmol/L HEPES (pH 7.4), 5 mmol/L EDTA, 2 mmol/L dithiothreitol, and 10 μmol/L digitonin for 15 minutes on ice. The lysates were clarified by centrifugation (12,000g), and supernatants were incubated with 50 μmol/L Asp-Glu-Val-Asp-aminomethylcoumarin (DEVD-AMC; Enzyme Systems Products, Inc at 37°C in the dark. Relative activities were then measured in a spectrofluorimeter (400 nm excitation, 505 nm emission); blanks included supernatants processed as outlined above without dye and supernatants preincubated with BDcmk (25 μmol/L).

Mitochondrial membrane potential measurements.

The potential-sensitive fluorochrome JC-1 (Molecular Probes, Eugene, OR) was used to measure ΔΨmito. Cells were obtained by centrifugation and incubated with 10 μmol/L JC-1 for 15 minutes at 37°C in the dark. Cells were washed in PBS and analyzed by FACS on the FL-2 channel (FACScan; Becton Dickinson, Mountain View, CA).

Annexin V binding.

Exposure of surface phosphatidylserine was quantified by surface annexin V staining as described previously.22 This assay was used as a DNA fragmentation-independent endpoint to confirm the involvement of apoptosis in the mechanism of cell death. Cells were resuspended in binding buffer containing 1 μg/mL FITC-conjugated annexin V (Nexins Research B.V., Hoeven, The Netherlands) and incubated for 30 minutes at 4°C, and cells were analyzed by flow cytometry (FACScan, Becton Dickinson).

Immunoblotting.

For detection of caspase-3, p53, p27, Fos, and Jun, cells were lysed for 1 hour at 4°C in a buffer containing 150 mmol/L NaCl, 1% Triton X-100, a cocktail of protease inhibitors (Complete Mini tablets; Boehringer-Mannheim, Indianapolis, IN), and 25 mmol/L Tris (pH 7.5). Debris was sedimented by centrifugation for 5 minutes at 12,000g, and the supernatants were solubilized for 5 minutes at 100°C in Laemmli’s SDS-PAGE sample buffer containing 100 mmol/L dithiothreitol.

For analysis of PARP cleavage samples were denatured in a urea/SDS buffer as follows: cells were incubated with various agents for 16 hours, obtained by centrifugation, and resuspended in 25 μL ice-cold PBS. Cells were subsequently disrupted by addition of 100 μL of a buffer containing 6 mol/L urea, 2% SDS, 10% glycerol, 5 mmol/L EDTA, 5% 2-mercaptoethanol, and 100 mmol/L Tris (pH 6.8) followed by pipeting through a 1-mL Pipetman tip, and samples were sonicated for 20 seconds at high power to sheer DNA. Samples were then incubated for 15 minutes at 65°C and centrifuged for 2 minutes at 12,000gbefore they were loaded onto 8% SDS-PAGE gels.

Polypeptides were resolved at 100 V on 8% to 12% gels and electrophoretically transferred to 0.2-μm nitrocellulose membranes (Schleicher & Schuell Inc) for 1 hour at 100 V. Membranes were blocked for 1 hour a TBS-T buffer (25 mmol/L Tris, pH 8.0, 150 mmol/L NaCl, and 0.05% Tween-20) containing 3% (wt/vol) nonfat dried milk. Blots were then probed overnight with primary antibody and developed using a horseradish peroxidase-coupled anti-mouse secondary antibody by enhanced chemiluminescence (Supersignal; Pierce Chemical Co, Rockford, IL) according to the manufacturer’s instructions.

Electrophoretic mobility shift (EMSA) assays.

Isolated nuclei were prepared by lysis with Triton X-100 and centrifugation through a glycerol cushion as described previously.23 Nuclear protein was extracted using a high salt, detergent-free buffer containing 20 mmol/L HEPES (pH 7.9), 400 mmol/L NaCl, 1 mmol/L EDTA, 1 mmol/L EGTA, 1 mmol/L dithiothreitol, and 1 mmol/L phenylmethylsulfonyl fluoride, for at least 20 minutes on ice. Extracts were centrifuged at 4°C for 5 minutes at 12,000g, and protein content in supernatants was measured by the Bradford method. A consensus double-stranded NFκB probe (obtained from Promega Inc, Madison, WI) was end-labeled using T4 polynucleotide kinase and γ-32P-ATP. Five to 20 μg of nuclear extract was then incubated in binding buffer (supplied by the manufacturer) containing 1 μg/mL poly dI:dC (Promega), ensuring that the final salt concentration was between 50 and 100 mmol/L. Reactions were incubated for 20 minutes at room temperature, and 1 μL of end-labeled probe was added. Samples were incubated for 30 minutes before addition of loading buffer (Promega) and electrophoresis on 4% nondenaturing polyacrylamide gels that were prerun in 0.5×Tris-borate-EDTA (TBE) buffer for 30 minutes at 100 V before use. Gels were run at 100 V in 0.5× TBE and dried, and DNA-protein complexes were detected by autoradiography.

Statistical analyses.

Mean values and standard deviations were calculated with Microsoft Excel (Microsoft Inc, Redmond, WA). Significance was evaluated using two-tailed paired Student’s t-tests with SPSS software (SPSS Inc, Chicago, IL).

Effects of proteasome inhibitor on DNA fragmentation and surface exposure of phosphatidylserine.

In a previous study we showed that a specific inhibitor of the NS protease completely suppressed apoptosis-associated DNA fragmentation in CLL cells treated with glucocorticoid or the nucleoside analog, fludarabine.9 Because proteasome inhibitors block apoptosis in thymocytes and neuronal cells,12,13 and another group has reported that the NS protease is homologous to the proteasome,10 we tested the effects of proteasome inhibitors on DNA fragmentation, measured by PI staining and FACS analysis, in CLL cells to determine whether the effects of zAPFcmk might be attributed to the proteasome. Levels were compared with those observed in response to treatment with glucocorticoid hormone. Our patient isolates fell into three general catetories: (1) those exhibiting relatively high (mean = 40%, n = 10) levels of apoptosis upon in vitro culture in the absence of hormone (“spontaneous”); (2) those exhibiting low spontaneous apoptosis but strong (mean = 40%, n = 28) increases in DNA fragmentation in response to glucocorticoid treatment (“sensitive”); and (3) those exhibiting low spontaneous apoptosis and low levels of glucocorticoid-induced DNA fragmentation (mean = 10%, n = 21) (“resistant”; Table 1). Strikingly, and contrary to our expectations, treatment with MG-132, a peptide-based proteasome antagonist, promoted high levels of DNA fragmentation in all three categories of cells (Table 1). Proteasome inhibitors were effective in all patient isolates analyzed (n = 59). Similar results were obtained with another, structurally distinct proteasome inhibitor, lactacystin (data not shown). Proteasome inhibitors also induced surface phosphatidylserine exposure, another downstream event in apoptosis that is thought to be independent of endonuclease activation (Fig 1B). Importantly, preincubation with zAPFcmk blocked MG-132–induced DNA fragmentation (Fig 1A), indicating that these inhibitors exert their effects on different biochemical activities.

Table 1.

Effects of Proteasome Inhibitor on DNA Fragmentation in CLL Patient Isolates

Patient Number Control*MPS MG132
Spontaneously apoptotic isolates
 
 
 156 46.1  60.5  87.3  
 160  27.4  34.1  82.7 
 180  89.8  90.1  96.1  
 213  24.9  48.5 70.1  
 227  37.9  49.8  93.8  
 228  51.2 71.2  97.3  
 230  32.4  57.7  90.4  
 232 25.6  37.8  91.2  
 233  29.5  31.8  72.8 
 235  38.5  57.6  89.9 
 Mean (SD)  40 (19) 54 (18)  87 (9) 
Glucocorticoidsensitive isolates  
 131 0.7  66.3  99.5  
 132  3.7  51  95.4 
 134  8.9  23.4  99.6  
 135  15.7  90.6 88.8  
 137  11.2  37.9  88.6  
 140  2.4 24.3  90.4  
 144  4.6  34.1  50.9  
 146 13.4  31  78.5  
 147  1.4  19.6  92.5 
 148  10  49.4  95.2  
 149  1.6  30.1 99  
 154  4.1  99.4  98.3  
 155  16.7 36.9  99.8  
 157  8.7  33.3  55.3  
 188 3.3  33.3  36.8  
 209  8  34.8  95.7 
 210  3.7  37.1  88.9  
 211  8.7  48.9 98.6  
 212  13.8  41.8  89.3  
 214  3.1 20.4  92.9  
 218  2.3  23.5  96.7  
 220 7.6  30.3  95  
 222  2.5  30.5  80.1 
 229  10.7  44.9  77.2  
 238  5.2  31.7 92.2  
 247  3  44.8  95.7  
 250  18.5 33.2  82.7  
 251  16.1  61.6  82.9 
 Mean (SD)  7 (5) 41 (19)  87 (15) 
Glucocorticoidresistant isolates  
 128 3.7  11  73.6  
 129  4.3  3.3  33.3 
 130  4.1  10.8  89  
 136  4.8  16.8 89.4  
 138  19.3  16.3  59.8  
 141  5.2 13  96.9  
 143  0.9  0.7  36.3  
 164 0.7  7  74.7  
 165  0.7  1.8  64.2  
 187 1.8  4.3  55  
 215  12.8  22.2  90.6 
 223  6.8  9.9  87.7  
 240  8  6.9  83 
 241  3.2  19.4  94.3  
 242  1.3  3.1 82  
 243  0.9  7.8  75.2  
 244  1.9 15.9  94.5  
 245  3.1  8.6  45.6  
 248 2.3  3.4  75  
 249  0.9  3.7  67.8  
 253 7.2  17.3  90.5  
 Mean (SD) 4 (4)  10 (6)  74 (19) 
Patient Number Control*MPS MG132
Spontaneously apoptotic isolates
 
 
 156 46.1  60.5  87.3  
 160  27.4  34.1  82.7 
 180  89.8  90.1  96.1  
 213  24.9  48.5 70.1  
 227  37.9  49.8  93.8  
 228  51.2 71.2  97.3  
 230  32.4  57.7  90.4  
 232 25.6  37.8  91.2  
 233  29.5  31.8  72.8 
 235  38.5  57.6  89.9 
 Mean (SD)  40 (19) 54 (18)  87 (9) 
Glucocorticoidsensitive isolates  
 131 0.7  66.3  99.5  
 132  3.7  51  95.4 
 134  8.9  23.4  99.6  
 135  15.7  90.6 88.8  
 137  11.2  37.9  88.6  
 140  2.4 24.3  90.4  
 144  4.6  34.1  50.9  
 146 13.4  31  78.5  
 147  1.4  19.6  92.5 
 148  10  49.4  95.2  
 149  1.6  30.1 99  
 154  4.1  99.4  98.3  
 155  16.7 36.9  99.8  
 157  8.7  33.3  55.3  
 188 3.3  33.3  36.8  
 209  8  34.8  95.7 
 210  3.7  37.1  88.9  
 211  8.7  48.9 98.6  
 212  13.8  41.8  89.3  
 214  3.1 20.4  92.9  
 218  2.3  23.5  96.7  
 220 7.6  30.3  95  
 222  2.5  30.5  80.1 
 229  10.7  44.9  77.2  
 238  5.2  31.7 92.2  
 247  3  44.8  95.7  
 250  18.5 33.2  82.7  
 251  16.1  61.6  82.9 
 Mean (SD)  7 (5) 41 (19)  87 (15) 
Glucocorticoidresistant isolates  
 128 3.7  11  73.6  
 129  4.3  3.3  33.3 
 130  4.1  10.8  89  
 136  4.8  16.8 89.4  
 138  19.3  16.3  59.8  
 141  5.2 13  96.9  
 143  0.9  0.7  36.3  
 164 0.7  7  74.7  
 165  0.7  1.8  64.2  
 187 1.8  4.3  55  
 215  12.8  22.2  90.6 
 223  6.8  9.9  87.7  
 240  8  6.9  83 
 241  3.2  19.4  94.3  
 242  1.3  3.1 82  
 243  0.9  7.8  75.2  
 244  1.9 15.9  94.5  
 245  3.1  8.6  45.6  
 248 2.3  3.4  75  
 249  0.9  3.7  67.8  
 253 7.2  17.3  90.5  
 Mean (SD) 4 (4)  10 (6)  74 (19) 

Apoptosis was quantified by PI staining and FACS analysis after 16 hours of culture as described in Materials and Methods. Spontaneously apoptotic isolates were arbitrarily defined as those exhibiting greater than 20% DNA fragmentation in control (untreated) cultures. Glucocorticoid-sensitive isolates were arbitrarily defined as those exhibiting greater than 20% increase in DNA fragmentation relative to controls following exposure to steroid.

Abbreviation: MPS, methylprednisolone.

*

For spontaneously apoptotic isolates, 16-hour control was used.

Fig. 1.

Induction of apoptosis by proteasome inhibition in CLL patient isolates. (A) DNA fragmentation analysis. Cells from a representative glucocorticoid-sensitive patient (see Table 1) were preincubated in the presence of 25 μmol/L zAPFcmk or 200 μmol/L zVADfmk for 1 hour and then treated with 10 μmol/L MG132, and DNA fragmentation was measured at 16 hours by PI staining and FACS analysis. (B) Surface phosphatidylserine exposure. Cells were incubated in absence or presence of 10 μmol/L methylprednisolone or 10 μmol/L MG132 with or without 200 μmol/L zVADfmk, and surface PS exposure was quantitated by staining with annexin-FITC and measured by FACS analysis. Results characteristic of three independent experiments with different patient isolates.

Fig. 1.

Induction of apoptosis by proteasome inhibition in CLL patient isolates. (A) DNA fragmentation analysis. Cells from a representative glucocorticoid-sensitive patient (see Table 1) were preincubated in the presence of 25 μmol/L zAPFcmk or 200 μmol/L zVADfmk for 1 hour and then treated with 10 μmol/L MG132, and DNA fragmentation was measured at 16 hours by PI staining and FACS analysis. (B) Surface phosphatidylserine exposure. Cells were incubated in absence or presence of 10 μmol/L methylprednisolone or 10 μmol/L MG132 with or without 200 μmol/L zVADfmk, and surface PS exposure was quantitated by staining with annexin-FITC and measured by FACS analysis. Results characteristic of three independent experiments with different patient isolates.

Close modal
Effects of MG-132 on DNA fragmentation in normal hematopoietic cells.

In a preliminary attempt to determine whether MG-132’s proapoptotic effects were selective for CLL cells, we analyzed the effects of the proteasome antagonist on DNA fragmentation in normal peripheral blood lymphocytes and in G-CSF–mobilized CD34+CD45+hematopoietic progenitor cells. Normal lymphocytes were killed by the compound, but the kinetics of the response were markedly delayed and the maximal response shifted from 16 hours to 48 hours (Fig 2A). Surface staining with specific B- and T-cell markers indicated that MG-132 was substantially more toxic to normal T cells than to B cells (data not shown). In contrast to the attenuated responses of lymphocytes, mobilized stem cells were highly sensitive to MG-132 (Fig 2B). Thus, proteasome inhibitors are capable of inducing apoptosis in certain normal as well as transformed hematopoietic cells.

Fig. 2.

Effect of proteasome inhibitors on normal lymphocytes. (A) Effects on normal lymphocytes. Isolated peripheral blood lymphocytes from normal donors were incubated in the presence of 10 μmol/L MG132 or 10 μmol/L methylprednisolone with or without 25 μmol/L zAPFcmk for 16 hours, and apoptosis was assessed by PI staining and FACS analysis. Results of one experiment are representative of three independent replicates. (B) Effects on hematopoietic progenitor cells. G-CSF–mobilized CD34+CD45+ progenitor cells were incubated in the absence or presence of 25 μmol/L zAPFcmk or 10 μmol/L MG-132 and DNA fragmentation was measured after 16 hours by PI staining and FACS analysis.

Fig. 2.

Effect of proteasome inhibitors on normal lymphocytes. (A) Effects on normal lymphocytes. Isolated peripheral blood lymphocytes from normal donors were incubated in the presence of 10 μmol/L MG132 or 10 μmol/L methylprednisolone with or without 25 μmol/L zAPFcmk for 16 hours, and apoptosis was assessed by PI staining and FACS analysis. Results of one experiment are representative of three independent replicates. (B) Effects on hematopoietic progenitor cells. G-CSF–mobilized CD34+CD45+ progenitor cells were incubated in the absence or presence of 25 μmol/L zAPFcmk or 10 μmol/L MG-132 and DNA fragmentation was measured after 16 hours by PI staining and FACS analysis.

Close modal
Effects of proteasome inhibition on caspase activation.

Caspases are a family of cysteine proteases that are thought to act at the core of the apoptotic pathway. Our previous work9 and that of others24,25 has confirmed that caspases are required for drug-induced apoptosis in CLL cells. We therefore investigated whether or not caspases were also required for apoptosis induced by proteasome inhibitors by four independent approaches. First, induction of apoptosis by proteasome inhibitors was also associated with specific cleavage of the caspase substrate, PARP, as detected by immunoblotting (Fig 3A). This occurred in all patient samples analyzed (n = 4), including one that did not respond to glucocorticoid treatment (Fig 3A). Second, proteasome inhibitors promoted hydrolysis of a specific caspase substrate (DEVD-AMC; Fig 3B). Third, proteasome inhibitors induced proteolytic processing of the inactive form of caspase-3 (procaspase-3; Fig 3C), providing more direct evidence for activation of caspase-3 and presumably other caspases in the response. Finally, the caspase inhibitor zVADfmk completely blocked MG-132–induced DNA fragmentation (Fig 1A) and surface exposure of phosphatidylserine (Fig 1B), confirming that caspase activation was required for proteasome inhibitor–induced apoptosis.

Fig. 3.

Caspase activation by proteasome inhibitors. (A) Cleavage of the caspase substrate, PARP, by treatment with proteasome inhibitor. Cells were treated with either 10 μmol/L methylprednisolone, 25 μmol/L zAPFcmk, or various doses of MG132. PARP was detected by immunoblotting. Intact (p116) and fragmented (p85) forms of PARP are indicated by arrows. (B) Effect of MG132 on DEVDase activity. Cells were treated in the absence or presence of 10 μmol/L MG132 for 8 hours, and hydrolysis of the caspase substrate DEVD-AMC was measured in a spectrofluorimeter. Results are from two experiments with independent patient isolates. Cells treated with 10 μmol/L BDcmk, a caspase inhibitor, did not show DEVDase activity more than 50 U above baseline levels. DNA fragmentation from these patients was measured in parallel. These patients are not included in Table 1. Patient no. 1: control = 18.0, MG132 = 96.9; patient no. 2: control = 1.4, MG132 = 32.5. (C) Activation of caspase-3 by proteasome inhibitor. Cells were treated with either 25 μmol/L zAPFcmk, 10 μmol/L MG132, or 10 μmol/L methylprednisolone for 16 hours and procaspase-3 was detected by immunoblotting. Results of one experiment representative of three replicates with independent patient isolates.

Fig. 3.

Caspase activation by proteasome inhibitors. (A) Cleavage of the caspase substrate, PARP, by treatment with proteasome inhibitor. Cells were treated with either 10 μmol/L methylprednisolone, 25 μmol/L zAPFcmk, or various doses of MG132. PARP was detected by immunoblotting. Intact (p116) and fragmented (p85) forms of PARP are indicated by arrows. (B) Effect of MG132 on DEVDase activity. Cells were treated in the absence or presence of 10 μmol/L MG132 for 8 hours, and hydrolysis of the caspase substrate DEVD-AMC was measured in a spectrofluorimeter. Results are from two experiments with independent patient isolates. Cells treated with 10 μmol/L BDcmk, a caspase inhibitor, did not show DEVDase activity more than 50 U above baseline levels. DNA fragmentation from these patients was measured in parallel. These patients are not included in Table 1. Patient no. 1: control = 18.0, MG132 = 96.9; patient no. 2: control = 1.4, MG132 = 32.5. (C) Activation of caspase-3 by proteasome inhibitor. Cells were treated with either 25 μmol/L zAPFcmk, 10 μmol/L MG132, or 10 μmol/L methylprednisolone for 16 hours and procaspase-3 was detected by immunoblotting. Results of one experiment representative of three replicates with independent patient isolates.

Close modal
Effects of proteasome inhibitor on mitochondrial function.

Disruption of mitochondria leading to the release of the electron transport chain intermediate, cytochrome c, has recently been implicated in caspase activation in other model systems.26Although the mechanisms underlying cytochrome c release are unclear, the event is associated with a drop in transmembrane potential (ΔΨ), which may facilitate the opening of transmembrane pores in the mitochondrial membrane that would allow passage of cytochrome c and other proapoptotic factors from the organelle.27 To determine whether this pathway of caspase activation was induced by proteasome inhibitors in CLL cells, we measured the effects of MG-132 on cytochrome c release in digitonin-permeabilized cells. MG-132 promoted rapid release of cytochrome c from mitochondria in intact CLL cells (Fig 4, lane 3), effects that were also observed in cells treated with zAPFcmk (Fig 4, lane 4) and to a lesser extent with glucocorticoid (Fig 4, lane 5). At this time point, control cells did not exhibit either increased levels of cytosolic cytochrome c (Fig4, compare lanes 1 and 2) or significant caspase activation (Fig 3B). The effects of zAPFcmk on cytochrome c release are consistent with earlier experments that showed that NS protease inhibition results in caspase activation9 (Fig 3). Proteasome inhibition also resulted in a drop in mitochondrial membrane potential, measured with the potential-sensitive dye, JC-1 (Fig 5). These results indicate that proteasome inhibitors promote caspase activation via direct or indirect effects on mitochondria.

Fig. 4.

Proteasome inhibition leads to release of cytochrome c from mitochondria. Cells were incubated in the absence (control) or presence of 10 μmol/L MG132, 25 μmol/L zAPFcmk, or 10 μmol/L methylprednisolone for 6 hours, and cytosolic cytochrome c was measured in digitonin-permeabilized cells by immunoblotting. Lane 1, 0 hours control; lane 2, 6 hours control; lane 3, MG132; lane 4, zAPFcmk; lane 5, methylprednisolone. Results are typical of three independent experiments with different CLL isolates.

Fig. 4.

Proteasome inhibition leads to release of cytochrome c from mitochondria. Cells were incubated in the absence (control) or presence of 10 μmol/L MG132, 25 μmol/L zAPFcmk, or 10 μmol/L methylprednisolone for 6 hours, and cytosolic cytochrome c was measured in digitonin-permeabilized cells by immunoblotting. Lane 1, 0 hours control; lane 2, 6 hours control; lane 3, MG132; lane 4, zAPFcmk; lane 5, methylprednisolone. Results are typical of three independent experiments with different CLL isolates.

Close modal
Fig. 5.

Effects of proteasome inhibition on mitochondrial membrane potential. Cells were incubated in the absence or presence of 10 μmol/L methylprednisolone with or without 25 μmol/L zAPFcmk or 10 μmol/L MG132. Mitochondrial membrane potential was assessed by the potential sensitive fluorochrome JC-1 and quantitated by FACS analysis.

Fig. 5.

Effects of proteasome inhibition on mitochondrial membrane potential. Cells were incubated in the absence or presence of 10 μmol/L methylprednisolone with or without 25 μmol/L zAPFcmk or 10 μmol/L MG132. Mitochondrial membrane potential was assessed by the potential sensitive fluorochrome JC-1 and quantitated by FACS analysis.

Close modal

The proteasome is known to degrade many proteins implicated in the control of cell survival, including p53,28Fos,29 Jun,30 Myc,31p27,32 and IκBα33 (a protein inhibitor of the transcription factor NFκB34). We did not observe any detectable alteration in the levels of p53, Fos, Jun, Myc, or p27 in CLL cells treated with MG-132 compared with cells incubated in medium alone (data not shown), suggesting that these polypeptides may not participate in proteasome inhibitor-induced apoptosis in this system. On the other hand, quantification of NFκB activity by EMSA revealed that both zAPFcmk and MG-132 drastically reduced the levels of active NFκB in isolated nuclei from CLL cells (Fig 6). Thus, blockade of the NFκB survival pathway may be responsible for triggering the disruption of mitochondrial function and caspase activation in these cells.

Fig. 6.

Inhibition of NFκB activity by NS protease and proteasome inhibitors. Cells were incubated for 6 hours in the absence (control) or presence of 25 μmol/L zAPFcmk or 10 μmol/L MG132, and NFκB activity was measured in isolated nuclear extracts by EMSA using an NFκB consensus element DNA probe. Lane 1, control extracts with excess unlabeled probe (specificity control); lane 2, 6 hours control; lane 3, zAPFcmk; lane 4, MG-132. Results of one experiment typical of over 20 independent replicates.

Fig. 6.

Inhibition of NFκB activity by NS protease and proteasome inhibitors. Cells were incubated for 6 hours in the absence (control) or presence of 25 μmol/L zAPFcmk or 10 μmol/L MG132, and NFκB activity was measured in isolated nuclear extracts by EMSA using an NFκB consensus element DNA probe. Lane 1, control extracts with excess unlabeled probe (specificity control); lane 2, 6 hours control; lane 3, zAPFcmk; lane 4, MG-132. Results of one experiment typical of over 20 independent replicates.

Close modal

In spite of the development of nucleoside analogs (fludarabine and cladribine) that have led to much better management of disease burden in CLL patients, CLL cells ultimately develop resistance to all currently available therapies, possibly because of apoptosis suppression. Our data show that the proteasome controls a central step in the maintenance of cell survival in CLL cells, such that inhibitors are capable of inducing apoptosis in all of them. The results support and extend independent work recently published by another group, who reported that the proteasome inhibitor lactacystin can promote radiation- and tumor necrosis factor (TNF)-induced apoptosis in CLL cells.35 The mechanism underlying the responses involves mitochondrial alterations leading to the release of cytochrome c and loss of mitochondrial membrane potential. The mitochondrial alterations are associated with caspase protease activation, as measured by specific cleavage of hallmark endogenous (PARP) and exogenous (DEVD-AMC) caspase substrates and proteolytic processing of procaspase-3. It is encouraging that we were not able to identify a single patient isolate exhibiting de novo resistance to proteasome inhibition among a fairly large (n = 59) panel, some of which (n = 21) showed marked resistance to glucocorticoid-induced apoptosis. However, our work does not address the issue of whether or not CLL cells can develop resistance to these agents under other conditions. Elimination of proteasome function in yeast results in lethality,36 but recent work suggests that mammalian cells contain another protease complex that can compensate for loss of proteasome function in cells chronically exposed to proteasome inhibitors.37 

Although the precise mechanism(s) precipitating the mitochondrial changes await further investigation, our preliminary efforts indicate that the effects of MG-132 are tightly linked to suppression of NFκB activity and not to stabilization of several other proteasome-regulated factors (p53, Fos, Jun, p27) that have been implicated in the control of apoptosis in other systems. Independent results obtained recently by another group support the idea that suppression of NFκB leads to apoptosis in CLL.35 The principal mechanism regulating NFκB activation involves an inhibitor protein, IκBα, that binds to NFκB and prevents its translocation to the nucleus.38Stimulation of cells with NFκB-activating signals results in phosphorylation of IκBα and its coupling to ubiquitin,33 a small (8 kD) polypeptide that forms polymers that serve to target proteins for destruction by the proteasome.39 In other B-cell model systems, constitutive NFκB activity is essential for cell survival40 and inhibition of NFκB using protease inhibitors or mutant forms of IκBα that cannot be degraded by the proteasome facilitates apoptosis in a number of different cell types.15-17Interestingly, the immunosuppressive effects of glucocorticoid hormones are linked to an inhibition of NFκB activity,41-44suggesting that suppression of NFκB may also be required for glucocorticoid-induced apoptosis. Our ongoing efforts are focused on further characterizing the role of NFκB in the maintenance of survival in CLL.

Even though some studies had shown that proteasome inhibitors trigger apoptosis in tumor cell lines,45,46 the observation that proteasome inhibition results in apoptosis in CLL was surprising to us. Proteasome inhibitors block glucocorticoid-induced apoptosis in immature thymocytes12 and the death of neuronal cells deprived of neurotrophins,13 and ubiquitin-dependent pathways appear involved in developmentally regulated cell death in the hawkmoth Manduca sexta and in radiation-induced apoptosis in tumor cells.47 Furthermore, we had shown that inhibitors of the NS protease, a putative proteasome homolog, completely block DNA fragmentation in CLL isolates.9 However, NS protease inhibitors did promote several other features of apoptosis, including caspase activation, mitochondrial dysregulation, and PS exposure, suggesting that their effects did overlap.9 Interestingly, a peptide inhibitor of the protease complex that can compensate for loss of proteasome function (AAFcmk)37 is very similar in sequence to our NS protease inhibitor (APFcmk), both of which contain a phenylalanine (F) residue at the critical P1 position. In addition, although MG132 failed to promote substantial accumulation of p53 in CLL cells, zAPFcmk consistently did (D.J. McConkey, unpublished observations, April 1998), and the effects of the inhibitors on NFκB activity are similar (Fig 6). We are currently isolating the NS protease complex, and a detailed analysis of its structure and biochemical regulation will reveal how it is related (if at all) to the proteasome.

The proteasome is central to normal cell physiology and hence complete inhibition of its activity is ultimately cytotoxic. However, appropriate titration of proteasome activity can elicit significant efficacy with limited side effects (Peter Elliot, Julian Adams, Proscript Inc, personal communication, April 1998). Indeed, the present report shows that proteasome inhibitors induce marked apoptosis in mobilized hematopoietic progenitor cells with more modest effects on normal lymphocytes. Moreover, extensive preclinical profiling of such compounds has clearly shown that maximum tolerated doses have only modest myelosuppressive activity (P. Elliot, J. Adams, personal communication). It is likely that progenitor cell mobilization and probably other manipulations that induce cell cycling will in general sensitize cells to proteasome inhibitor-induced apoptosis, as previous work suggests that proliferating cells are more sensitive to their effects than are postmitotic cells.48 Additionally, the proteasome inhibitor, PS-341, has been shown to possess antitumor activity of its own and this effect is enhanced when combined with other chemotherapeutics (P. Elliot, J. Adams, personal communication). With the advent of acceptable phase I safety data, the present results would strongly argue that PS-341 should be evaluated in patients with refractory CLL.

The authors thank Virginia Snell for providing the purified hematopoietic stem cells, Yuko Miyamoto for purified peripheral lymphocytes, and Julian Adams and Peter Elliot (Proscript Inc, Cambridge, MA) for sharing preliminary data on PS-341.

Supported by grants from the National Institutes of Health (CA16672, CA55164, CA49639) (to M.A.), Physicians’ Referral Service, MDACC, the American Cancer Society (RPG-97-169-01-CDD) (to D.J.M.), and an American Legion Auxiliary Fellowship (to J.C.).

The publication costs of this article were defrayed in part by page charge payment. This article must therefore be hereby marked “advertisement” in accordance with 18 U.S.C. section 1734 solely to indicate this fact.

1
Robertson
LE
Chubb
S
Meyn
RE
Story
M
Ford
R
Hittelman
WN
Plunkett
W
Induction of apoptotic cell death in chronic lymphocytic leukemia by 2-chloro-2’-deoxyadenosine and 9-β-D-arabinosyl-2’-fluoroadenine.
Blood
81
1993
143
2
Consoli
U
El-Tounsi
I
Sandoval
A
Snell
V
Kleine
HD
Brown
W
Robinson
JR
DiRaimondo
F
Plunkett
W
Andreeff
M
Differential induction of apoptosis by fludarabine monophoshate in leukemic B and normal T cells in chronic lymphocytic leukemia.
Blood
91
1998
1742
3
Huang
P
Robertson
LE
Wright
S
Plunkett
W
High molecular weight DNA fragmentation: A critical event in nucleoside analogue-induced apoptosis in leukemia cells.
Clin Cancer Res
1
1995
1005
4
Henkart
PA
ICE family proteases: Mediators of all apoptotic cell death?
Immunity
4
1996
195
5
Martin
SJ
Green
DR
Protease activation during apoptosis: Death by a thousand cuts?
Cell
82
1995
349
6
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
7
Liu
X
Kim
CN
Yang
J
Jemmerson
R
Wang
X
Induction of the apoptotic program in cell-free extracts: Requirement for dATP and cytochrome c.
Cell
86
1996
147
8
Cohen
GM
Caspases: The executioners of apoptosis.
Biochem J
326
1997
1
9
Chandra
J
Gilbreath
J
Freireich
EJ
Kliche
KO
Andreeff
M
Keating
M
McConkey
DJ
Protease activation is required for glucocorticoid-induced apoptosis in chronic lymphocytic leukemic lymphocytes.
Blood
90
1997
3673
10
Clawson
GA
Norbeck
LL
Hatem
CL
Rhodes
C
Amiri
P
McKerrow
JH
Patierno
SR
Fiskum
G
Ca2+-regulated serine protease associated with the nuclear scaffold.
Cell Growth Differ
3
1992
827
11
Schwartz
LM
Smith
SW
Jones
MEE
Osborne
BA
Do all programmed cell deaths occur via apoptosis?
Proc Natl Acad Sci USA
90
1993
980
12
Grimm
LM
Goldberg
AL
Poirier
GG
Schwartz
LM
Osborne
BA
Proteasomes play an essential role in thymocyte apoptosis.
EMBO J
15
1996
3835
13
Sadoul
R
Fernandez
PA
Quiquerez
AL
Martinou
I
Maki
M
Schroter
M
Becherer
JD
Irmler
M
Tschopp
J
Martinou
JC
Involvement of the proteasome in the programmed cell death of NGF-deprived sympathetic neurons.
EMBO J
15
1996
3845
14
Beg
AA
Sha
WC
Bronson
RT
Ghosh
S
Baltimore
D
Embryonic lethality and liver degeneration in mice lacking the RelA component of NFκB.
Nature
376
1995
167
15
Beg
AA
Baltimore
D
An essential role for NFκB in preventing TNFα-induced cell death.
Nature
274
1996
782
16
Wang
CY
Mayo
MW
Baldwin
AS
TNF- and cancer therapy-induced apoptosis: Potentiation by inhibition of NFκB.
Nature
274
1996
784
17
Antwerp
DJV
Martin
SJ
Kafri
T
Green
DR
Verma
IM
Suppression of TNFa-induced apoptosis by NFκB.
Nature
274
1996
787
18
Rai
KR
Sawitsky
A
Cronkite
EP
Chanana
AD
Levy
RN
Pasternack
BS
Clinical staging of chronic lymphocytic leukemia.
Blood
46
1975
219
19
Sutherland
DR
Keating
A
Nayar
R
Anania
S
Stewart
AK
Sensitive detection and enumeration of CD34+ cells in peripheral and cord blood by flow cytometry.
Exp Hematol
22
1994
1003
20
Nicoletti
I
Migliorati
G
Pagliacci
MC
Grignani
F
Riccardi
C
A rapid and simple method for measuring thymocyte apoptosis by propidium iodide staining and flow cytometry.
J Immunol Methods
139
1991
271
21
Yang
J
Liu
X
Bhalla
K
Kim
CN
Ibrado
AM
Cai
J
Peng
TI
Jones
DP
Wang
X
Prevention of apoptosis by bcl-2: Release of cytochrome c from mitochondria blocked.
Science
275
1997
1129
22
Koopman
G
Reutelingsperger
CPM
Kuijten
GAM
Keelman
RMJ
Pals
ST
Oers
MHJv
Annexin V for flow cytometric detection of phosphatidylserine expression on B cells undergoing apoptosis.
Blood
84
1994
1415
23
McConkey
DJ
Chandra
J
Wright
S
Plunkett
W
McDonnell
TJ
Reed
JC
Keating
MJ
Apoptosis sensitivity in chronic lymphocytic leukemia is determined by endogenous endonuclease content and relative expression of BCL-2 and BAX.
J Immunol
156
1996
2624
24
Bellosillo
BDM
Colomer
D
Gil
J
Involvement of CED-3/ICE proteases in the apoptosis of B-chronic lymphocytic leukemia cells.
Blood
89
1997
3378
25
Krajewski
S GR
Zapata
JM
Krajewska
M
Kitada
S
Chhanabhai
M
Horsman
D
Berean
K
Piro
LD
Fugier-Vivier
I
Liu
YJ
Wang
HG
Reed
JC
Immunolocalization of the ICE/Ced-3-family protease, CPP-32 (Caspase-3), in non-Hodgkin’s lymphomas, chronic lymphocytic leukemias, and reactive lymph nodes.
Blood
89
1997
3817
26
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
27
Kroemer
G
Zamzami
N
Susin
SA
Mitochondrial control of apoptosis.
Immunol Today
18
1997
44
28
Lopes
UG
Erhardt
P
Yao
R
Cooper
GM
p53-dependent induction of apoptosis by proteasome inhibitors.
J Biol Chem
272
1997
12893
29
Stancovski
I
Gonen
H
Orian
A
Schwartz
AL
Ciechanover
A
Degradation of the proto-oncogene product c-Fos by the ubiquitin proteolytic system in vivo and in vitro: Identification and characterization of the conjugating enzymes.
Mol Cell Biol
15
1995
7106
30
Jariel-Encontre
I
Pariat
M
Martin
F
Carillo
S
Salvat
C
Piechaczyk
M
Ubiquitinylation is not an absolute requirement for degradation of c-Jun protein by the 26 S proteasome.
J Biol Chem
270
1995
11623
31
Bonvini
P
Nguyen
P
Trepel
J
Neckers
LM
In vivo degradation of N-myc in neuroblastoma cells is mediated by the 26S proteasome.
Oncogene
16
1998
1131
32
Pagano
M
Tam
SW
Theodoras
AM
Beer-Romero
P
Sal
GD
Chau
V
Yew
PR
Draetta
GF
Rolfe
M
Role of the ubiquitin-proteasome pathway in regulating abundance of the cyclin-dependent kinase inhibitor p27.
Science
269
1995
682
33
Traenckner
EBM
Wilk
S
Baeuerle
PA
A proteasome inhibitor prevents activation of NF-κB and stabilizes a newly phosphorylated form of IkB-α that is still bound to NF-κB.
EMBO J
13
1994
5433
34
Verma
IM
Stevenson
JK
Schwartz
EM
Antwerp
DV
Miyamoto
S
Rel/NFκB/IκB family: Intimate tales of association and dissociation.
Genes Dev
9
1996
2723
35
Delic
J
Masdehors
P
Omura
S
Cosset
JM
Dumont
J
Binet
JL
Magdelenat
H
The proteasome inhibitor lactacystin induces apoptosis and sensitizes chemo- and radioresistant human chronic lymphocytic leukaemic lymphocytes to TNF alpha-initiated apoptosis.
Br J Cancer
77
1998
1103
36
Ghislain
M
Udvardy
A
Mann
C
S. cerevisiae 26S protease mutants arrest cell division in G2/metaphase.
Nature
366
1993
358
37
Glas
R
Bogyo
M
McMaster
JS
Gaczynska
M
Ploegh
HL
A proteolytic system that compensates for loss of proteasome function.
Nature
392
1998
618
38
Baeuerle
PA
Baltimore
D
NF-κB: Ten Years After.
Cell
87
1996
13
39
Hochstrasser
M
Ubiquitin, proteasomes, and the regulation of intracellar protein degradation.
Curr Opin Cell Biol
7
1995
215
40
Wu
M
Lee
H
Bellas
RE
Schauer
SL
Arsura
M
Katz
D
FitzGerald
MJ
Rothstein
TL
Sherr
DH
Sonenshein
GE
Inhibition of NFκB/Rel induces apoptosis of murine B cells.
EMBO J
15
1995
4682
41
Ray
A
Prefontaine
KE
Physical association and functional antagonism between the p65 subunit of transcription factor NFκB and the glucocorticoid receptor.
Proc Natl Acad Sci USA
91
1994
752
42
Scheinman
RI
Gualberto
A
Jewell
CM
Cidlowski
JA
Baldwin
AS
Characterization of mechanisms involved in transrepression of NFκB by activated glucocorticoid receptors.
Mol Cell Biol
15
1995
943
43
Scheinman
RI
Cogswell
PC
Lofquist
AK
Baldwin
AS
Role of transcription activation of IκBα in mediation of immunosuppression by glucocorticoids.
Science
270
1995
283
44
Auphan
N
DiDonato
JA
Rosette
C
Helmberg
A
Karin
M
Immunosuppression by glucocorticoids: Inhibition of NFκB activity through induction of IκBα synthesis.
Science
270
1995
286
45
Drexler
HCA
Activation of the cell death program by inhibition of proteasome function.
Proc Natl Acad Sci USA
94
1997
855
46
Soldatenkov
VA
Dritschilo
A
Apoptosis of Ewing’s sarcoma cells is accompanied by accumulation of ubiquinated proteins.
Cancer Res
57
1997
3881
47
Delic
J
Morange
M
Magdelenat
H
Ubiquitin pathway involvement in human lymphocyte γ-irradiation-induced apoptosis.
Mol Cell Biol
13
1993
4875
48
Drexler
HCA
Programmed cell death and the proteasome.
Apoptosis
3
1998
1

Author notes

Address reprint requests to David J. McConkey, PhD, Department of Cell Biology - 173, U.T. M.D. Anderson Cancer Center, 1515 Holcombe Blvd, Houston, TX 77030; email: dmcconke@notes.mdacc.tmc.edu.

Sign in via your Institution