A-1155463

Exploiting selective BCL-2 family inhibitors to dissect cell survival dependencies and define improved strategies for cancer therapy
Joel D. Leverson,1* Darren C. Phillips,1 Michael J. Mitten,1 Erwin R. Boghaert,1 Dolores Diaz,2 Stephen K. Tahir,1 Lisa D. Belmont,2 Paul Nimmer,1 Yu Xiao,1 Xiaoju Max Ma,2† Kym N. Lowes,3,4 Peter Kovar,1 Jun Chen,1 Sha Jin,1 Morey Smith,1 John Xue,1 Haichao Zhang,1 Anatol Oleksijew,1 Terrance J. Magoc,1 Kedar S. Vaidya,1 Daniel H. Albert,1 Jacqueline M. Tarrant,2 Nghi La,2
Le Wang,1 Zhi-Fu Tao,1 Michael D. Wendt,1 Deepak Sampath,2 Saul H. Rosenberg,1 Chris Tse,1 David C. S. Huang,3,4 Wayne J. Fairbrother,2 Steven W. Elmore,1 Andrew J. Souers1

The BCL-2/BCL-XL/BCL-W inhibitor ABT-263 (navitoclax) has shown promising clinical activity in lymphoid malignan- cies such as chronic lymphocytic leukemia. However, its efficacy in these settings is limited by thrombocytopenia caused by BCL-XL inhibition. This prompted the generation of the BCL-2–selective inhibitor venetoclax (ABT-199/GDC- 0199), which demonstrates robust activity in these cancers but spares platelets. Navitoclax has also been shown to enhance the efficacy of docetaxel in preclinical models of solid tumors, but clinical use of this combination has been limited by neutropenia. We used venetoclax and the BCL-XL–selective inhibitors A-1155463 and A-1331852 to assess the relative contributions of inhibiting BCL-2 or BCL-XL to the efficacy and toxicity of the navitoclax-docetaxel combination. Selective BCL-2 inhibition suppressed granulopoiesis in vitro and in vivo, potentially accounting for the exacerbated neutropenia observed when navitoclax was combined with docetaxel clinically. By contrast, selec- tively inhibiting BCL-XL did not suppress granulopoiesis but was highly efficacious in combination with docetaxel when tested against a range of solid tumors. Therefore, BCL-XL–selective inhibitors have the potential to enhance the efficacy of docetaxel in solid tumors and avoid the exacerbation of neutropenia observed with navitoclax. These studies demonstrate the translational utility of this toolkit of selective BCL-2 family inhibitors and highlight their potential as improved cancer therapeutics.

 

INTRODUCTION
Multicellular organisms have evolved molecular mechanisms to elim- inate cells that are no longer required or that have become compro- mised through environmental insults. This active process of programmed cell death, or apoptosis, is especially important for eliminating cells with the potential for malignant transformation. The ability to suppress apoptosis under conditions of cell stress has been defined as one of the hallmarks of a cancer cell (1), and thus, triggering apoptosis in can- cer cells represents a powerful therapeutic approach.
Apoptosis is regulated by a family of closely related proteins exem- plified by BCL-2 (B cell lymphoma protein 2), the first family member discovered. BCL-2 family proteins are defined by one to four BCL-2 homology motifs (BH1 to BH4) and can be subdivided into pro- and antiapoptotic subsets. Proapoptotic proteins include the BH3-only pro- teins such as BIM, BAD, BID, and NOXA, and the BH1 to BH4 pro- teins BAK and BAX, which serve as the ultimate effectors of apoptosis by oligomerizing to form pores in the mitochondrial outer membrane. BCL-2 and the closely related antiapoptotic proteins BCL-XL and MCL-1 can sequester their proapoptotic counterparts by binding to their BH3 motifs and thereby inhibit the initiating steps of programmed cell death (2–4).
Small molecules that mimic the BH3 motif have been developed with the aim of binding antiapoptotic proteins, liberating proapoptotic proteins, and triggering apoptosis in cancer cells, many of which are thought to be “primed for death” due to the expression of high levels of pro- and antiapoptotic protein complexes (5). ABT-737 and the re- lated orally bioavailable compound ABT-263 (navitoclax) are BH3 mi- metics that bind with high affinity to BCL-2, BCL-XL, and BCL-W, but not to MCL-1 (6, 7). Preclinically, navitoclax inhibits tumor growth as a single agent (8) and in combination with standard-of-care therapeu- tics such as gemcitabine, vincristine, and docetaxel (9, 10). In clinical studies, navitoclax exhibited objective antitumor activity in lymphoid malig- nancies but also induced rapid, reversible, and dose-dependent throm- bocytopenia that was dose-limiting in this setting (11, 12). Navitoclax-induced thrombocytopenia was found to be a consequence of inhibiting BCL-XL (13, 14), and this prompted the development of ABT-199/GDC-0199 (venetoclax), a BCL-2–selective inhibitor that maintains efficacy in hema- tologic tumor models but spares platelets (15). Venetoclax has shown promising signs of clinical activity in hematologic malignancies, demon- strating objective responses in chronic lymphocytic leukemia (CLL) (16) and non-Hodgkin’s lymphoma (17).
Navitoclax has also been evaluated for the treatment of solid tumors in combination with chemotherapeutics. When combined with docetax-

1AbbVie Inc., North Chicago, IL 60064, USA. 2Genentech Inc., South San Francisco, CA 94080, USA. 3Walter and Eliza Hall Institute of Medical Research, Parkville, Victoria 3052, Australia. 4Department of Medical Biology, University of Melbourne, Parkville, Victoria 3010, Australia. *Corresponding author. E-mail: [email protected]
†Present address: Department of Genomics and Oncology, Roche Molecular Systems Inc., 4300 Hacienda Drive, Pleasanton, CA 94588, USA.
el, febrile neutropenia was the most commonly observed dose-limiting toxicity for navitoclax (18). However, it has remained unclear whether this effect is driven by the inhibition of BCL-2, BCL-XL, or both proteins. We reasoned that, if the contributions of BCL-XL and BCL-2 inhibition could be defined, an improved strategy to minimize toxicity and thus
maximize antitumor activity might be identified. To this end, we gen- erated the BCL-XL–selective inhibitors A-1155463 (19) and A-1331852 to complement the BCL-2–selective agent venetoclax and its close an- alog A-1211212. Equipped with this toolkit, we now had the opportu- nity to chemically dissect the biological activities of navitoclax and attribute them to the inhibition of BCL-2, BCL-XL, or both targets.
In the studies described here, we use these inhibitors to define the roles of BCL-2 and BCL-XL in maintaining the survival of multiple hema- tologic and solid tumor cell lines. Using a variety of in vitro, ex vivo, and in vivo model systems, we also parse the contributions of BCL-2 and BCL-XL inhibition to the antitumor efficacy and the myelotoxicity observed when navitoclax is combined with docetaxel. Our data indicate that BCL-XL inhibition is sufficient to recapitulate the efficacy of navitoclax in combination with docetaxel, whereas BCL-2 inhibition likely accounts for navitoclax-related neutropenia. Just as venetoclax rep- resented an improvement over the efficacy/toxicity profile of navitoclax for hematologic malignancies, so BCL-XL–selective inhibitors may be better tolerated, and thus enable greater efficacy, when combined with chemotherapeutic agents for the treatment of solid tumors.
RESULTS

Selective BCL-2 family inhibitors parse the activity of navitoclax into BCL-2– and BCL-XL–dependent components Navitoclax most closely mimics the BH3-only protein BAD, binding to BCL-2 (Ki = 0.044 nM) and BCL-XL (Ki = 0.055 nM) with high affinity and more weakly to BCL-W (Ki = 21 nM; Table 1). Because its affinity for BCL-W is >300-fold weaker, the biological effects of navitoclax are most likely due to the inhibition of BCL-2, BCL-XL, or both of these targets (Fig. 1A). Whereas venetoclax enables one to probe specifically for BCL-2–dependent phenomena, a BCL-XL–selective coun- terpart was needed to parse the effects of navitoclax more complete- ly. We used the recently described BCL-XL–selective inhibitor WEHI-539 (20) as an effective lead structure to generate a more potent and chemi- cally stable molecule, A-1155463 (19) (Fig. 1A). A-1155463 binds to BCL-XL with high affinity (Ki < 0.010 nM) but has much weaker affinity for BCL-2 (Ki = 74 nM), BCL-W (Ki = 8 nM), and MCL-1 (Ki > 444 nM) (Table 1). As predicted by this binding profile, A-1155463 disrupts BCL-XL–BIM but not BCL-2–BIM complexes in cells (Fig. 1B). A-1155463 kills BCL-XL–dependent Molt-4 cells (EC50 = 70 nM) but

has no measurable cytotoxicity against BCL-2–dependent RS4;11 cells (EC50 > 5 mM) (Table 1). A-1155463 induces the hallmarks of apopto- sis, as evidenced by the release of cytochrome c from mitochondria, caspase activation, and the accumulation of caspase-dependent sub– G0-G1 DNA content in BCL-XL–dependent H146 cells (8, 15) (Fig. 1, C to E). In the absence of the essential apoptosis effector proteins BAK and BAX, no toxicity was observed with A-1155463 (fig. S1), thus exclud- ing any major off-target cytotoxicity. As anticipated (20–22), potent killing of murine embryonic fibroblasts (MEFs) was observed only when Mcl-1 was deleted (fig. S1), demonstrating that this compound acted in a highly specific manner. Together, venetoclax and A-1155463 provide the ability to separate the activity of navitoclax into its respective BCL-2– and BCL-XL–dependent components.
As a first implementation of this chemical toolkit, we investigated the roles of BCL-2 and BCL-XL in mediating the survival of cancer cell lines with known sensitivity to navitoclax. Small cell lung cancer (SCLC) cell lines have shown sensitivity to navitoclax that is inversely corre- lated with the expression of MCL-1 (23). Because BCL2 copy number gains were observed in many of these cell lines (24), it was reasonable to infer that this activity was due to the inhibition of BCL-2. Indeed, among a panel of 11 SCLC cell lines, two with BCL2 amplification, NCI-H889 and NCI-H211, were sensitive (EC50 ≤ 1.0 mM) to venetoclax but not to A-1155463 (Fig. 2A and Table 2), indicating that these cell lines were BCL-2–dependent. However, most navitoclax-sensitive SCLC lines either were more sensitive to the BCL-XL–selective inhibitor alone (NCI-H446, NCI-H847, NCI-H1417, and NCI-H1836) or required in- hibition of both BCL-2 and BCL-XL. For example, navitoclax killed NCI-H345, NCI-H69, DMS79, and NCI-H1048 more potently than did either selective inhibitor alone. Notably, the combination of venetoclax and A-1155463 was able to recapitulate the effect of navitoclax in this setting, exhibiting synergistic killing of NCI-H69 (Bliss sum = 730 ± 26) and NCI-H345 (Bliss sum = 603 ± 89) cells (Fig. 2B). NCI-H187 cells were susceptible to all three inhibitors, indicating that inhibition of either BCL-2 or BCL-XL alone was sufficient to trigger their killing. These data demonstrate the utility of venetoclax and A-1155463 for determining whether BCL-2 or BCL-XL is necessary and sufficient for the survival of a given cell line. Furthermore, they indicate that BCL-2 and BCL-XL do not function redundantly in all cellular contexts.
We performed similar parsing studies in a panel of 24 acute myeloid leukemia (AML) cell lines, 9 of which were sensitive to venetoclax (EC50 ≤ 0.1 mM) (Fig. 2C and Table 2), consistent with a recent report
Table 1. Biochemical and cellular activity of small-molecule BCL-2 family inhibitors. For each compound listed, binding affinities (Ki) for BCL-2 family proteins were calculated in time-resolved fluorescence resonance energy

transfer (TR-FRET) assays, and the effects on cell viability [median effective concentration (EC50)] were assessed in BCL-2–dependent (RS4;11) or BCL- XL–dependent (Molt-4) cancer cell lines. All values are in nanomolar units.
BCL-XL/BCL-2 BCL-2–selective BCL-XL–selective

Navitoclax A-874009 Venetoclax A-1211212 A-1155463 A-1331852

BCL-2 TR-FRET 0.044 0.048 <0.010 <0.010 74 6
BCL-XL TR-FRET 0.055 0.463 48 14 <0.010 <0.010
MCL-1 TR-FRET >224 >3900 >444 >444 >444 142
BCL-W TR-FRET 21 2 245 852 8 4
Molt-4 viability 303 134 >5000 >5000 70 6
RS4;11 viability 112 28 8 6 >5000 >5000

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Fig. 1. Selective BCL-2 family inhibitors enable the functional dissection of the effects of navitoclax. (A) Chemical structures and selectivity profiles of BH3 mimetics used for in vitro studies. ABT-263 (navitoclax) inhibits both BCL-2 and BCL-XL, whereas ABT-199 (venetoclax) selectively inhibits BCL-2 and A-1155463 se- lectively inhibits BCL-XL. (B) Quantitative measurement of BCL-XL–BIM and BCL-2–BIM complexes in Molt-4 cells after 4-hour treatments with increasing concentrations of A-1155463. Data represent the average of triplicate experiments, with error bars indicating the SD. (C) Cytochrome c levels present in NCI-H146 mito- chondrial and cytosolic fractions as determined by immunoblotting after treatments with increasing con- centrations of A-1155463, venetoclax, or navitoclax for 4 hours. (D) Caspase-3/7 activation in NCI-H146 cells after incubation with increasing concentrations of A-1155463 for 4 hours. Data represent the average of triplicate experiments, with error bars indicating the SD. (E) Apoptosis (sub–G0-G1 accumulation) as assessed by fluorescence-activated cell sorting in NCI-H146 cells after 1-hour preincubation plus or minus the caspase inhibitor Z-VAD-fmk (75 mM) and an additional 24 hours of incubation with increasing concentrations of A-1155463. Data represent the average of duplicate experiments, with error bars indicating the SD.

BCL-2/BCL-XL co-dependent SCLC cell lines, KG-1 (Bliss sum = 892 ± 22) and SKM-1 (Bliss sum = 997 ± 154) showed synergistic sensitivity to the combination of venetoclax and A-1155463 (Fig. 2D). OCI-AML3, NOMO-1, and ME-1 cells were resistant to all three inhibitors, potentially implicating other BCL-2 family members in mediating their survival. Indeed, MCL-1 was shown to maintain the survival of OCI-AML3 when treated with ABT-737 or venetoclax (25).

Synergistic killing of cancer cell lines by the navitoclax-docetaxel combination is driven by
BCL-XL inhibition
ABT-737 and navitoclax have been shown to synergize with taxanes in killing a vari- ety of cancer cell lines (10, 26–28). To dissect the contributions of BCL-2 and BCL-XL inhibition to the activity of this combination, we tested panels of breast cancer, non–small cell lung cancer (NSCLC), and ovarian can- cer cell lines with combinations of docetaxel and navitoclax, venetoclax, or A-1155463. To assess potential combination activity, we used the Bliss independence model (28–30), according to which negative integers indi- cate antagonism, a value of zero indicates additive activity, and positive integers in- dicate synergy. Bliss scores were calculated for each combination in the dose matrix and then totaled to give a “Bliss sum” (Table 3). For the purposes of these studies, Bliss sum values >150 were considered in- dicative of synergy. The navitoclax-docetaxel combination demonstrated synergistic kill- ing of several breast cancer cell lines (fig. S2) and enhanced the induction of pro- grammed cell death, as indicated by ele- vated caspase cleavage, poly(ADP-ribose) polymerase cleavage, and the accumulation of cells with sub–G0-G1 DNA content (fig. S3). The Bliss sum values for the A-1155463– docetaxel combinations showed a significant correlation with those of the navitoclax- docetaxel combinations when compared among breast cancer (r = 0.75, P < 0.0001, n = 28) (Fig. 3A) or NSCLC (r = 0.94, P < 0.0001, n = 15) (Fig. 3B) cell lines. However, venetoclax-docetaxel Bliss values showed no correlation to those calculated for navitoclax-

(25). Most AML cell lines (14 of 24) were more sensitive to venetoclax than to a BCL-XL–selective inhibitor. Cell lines bearing the JAK2 V617F mutation (UKE-1, SET-2, and HEL) were a notable exception, showing sensitivity to the BCL-XL–selective inhibitor but not to venetoclax. Again, the inhibition of both BCL-2 and BCL-XL was required for the effective killing of some cell lines, including KG-1 and SKM-1. Like
docetaxel. In addition, the navitoclax-docetaxel and A-1155463–docetaxel combinations exhibited synergistic killing of four of six ovarian cancer cell lines tested (Bliss sum > 150), but no synergy was observed with the venetoclax-docetaxel combination (fig. S4). Collectively, these data in- dicate that BCL-XL is the key target of navitoclax for inducing the syn- ergy observed with docetaxel in these solid tumor models.

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 
Fig. 2. Venetoclax and A-1155463 define the BCL-2 family dependence profile of cancer cell lines. (A) SCLC cell lines were incubated with increasing concentrations of navitoclax (BCL-2/BCL-XL), venetoclax (BCL-2–selective), or A-1155463 (BCL-XL–selective) for 48 hours before assessing cell viability. Cell killing EC50 values are plotted for each compound against the cell lines examined. (B) NCI-H69 and NCI-H345 cells were incubated with increasing concentrations of A-1155463 in the presence or absence of venetoclax for 48 hours before assessing cell viability. (C) AML cell lines were treated as in (A) before assessing cell viability. Cell killing EC50 values are plotted for each compound against the cell lines examined. (D) KG-1 and SKM-1 cells were incubated with increasing concentrations of A-1155463 in the presence or absence of venetoclax for 48 hours before assessing cell viability.

design to generate A-1331852 (Fig. 4A), a BCL-XL–selective inhibitor with oral bio- availability. A-1331852 is a potent BCL-XL inhibitor, binding BCL-XL with a Ki value of <0.010 nM and demonstrating cellular activity 10- to 50-fold more potent than A-1155463 and navitoclax, respectively (Table 1). This molecule selectively disrupts BCL-XL–BIM complexes and induces the hallmarks of apoptosis in BCL-XL–dependent Molt-4 cells with median inhibitory con- centration (IC50) values in the low nano- molar range (Fig. 4, B to E, and Table 1) but does not affect MEF cells lacking BAK or BAX (fig. S5). Moreover, A-1331852 dem- onstrates antitumor efficacy in the Molt-4 xenograft model, inducing tumor regres- sions as a single agent (Fig. 4F). Addition- ally, A-1331852 combines with venetoclax to recapitulate the efficacy of navitoclax in the NCI-H1963.FP5 xenograft model of SCLC (Fig. 4G), thus providing in vivo confirmation of the combination studies shown in Fig. 2 (B and D).
Inhibition of tumor growth by A-1331852 combined with docetaxel was determined in seven subcutaneous xenograft models of solid tumors, including breast cancer, NSCLC, and ovarian cancer. Given as a single agent, A-1331852 significantly (P < 0.05) inhibited tumor growth in all seven models (Table 4). Although its single-agent activity was modest (TGImax < 60% in five of seven models), A-1331852 increased the efficacy of docetaxel in all seven models. As shown in Table 4, the maximum tumor growth inhibition (TGImax) for A-1331852 as a single agent ranged between 34% (OVCAR-5) and 67% (A549-FP3). The most durable response to A-1331852 was a tumor growth delay (TGD) of 108%, observed in the A549-FP3 model. This indicates that the median time required for the tumors to reach a volume of 1 cm3 is about twice as long when treated with A-1331852 as compared to a sham-treated control. When comparing the combination to the most ef- fective single-agent treatment, the increase in amplitude and durability of the response was statistically significant (P < 0.05) in five of seven models. The effect was most pronounced in the MDA-MB-231 LC3 me- tastatic breast cancer model (Fig. 5A) and the NSCLC models NCI-H1650 (Fig. 5B)

Orally bioavailable BCL-XL–selective inhibitor A-1331852 enhances the efficacy of docetaxel in vivo
We next assessed the ability of a selective BCL-XL inhibitor to enhance the efficacy of docetaxel in vivo. To this end, we used structure-based
and NCI-H358 (Table 4). Overall, the single-agent and combination treat- ments were well tolerated by mice, without overt signs of toxicity or weight loss of >9%. These data demonstrate that BCL-XL inhibition alone can enhance the efficacy of docetaxel in a variety of solid tumor models.
Table 2. Cellkillingactivityofsmall-moleculeBCL-2familyinhibitors. SCLC and AML cell lines were incubated with increasing concentrations of navitoclax, venetoclax, or A-1155463 for 48 hours before assessing cell viability. Cell killing EC50 values are listed for each compound against the cell lines examined.

CellTiter-Glo cell viability assay

in the clinic (18). Using the toolkit of dual and selective inhibitors, we next asked if BCL-2 or BCL-XL inhibition alone might be sufficient to suppress granulopoiesis in colony-forming assays. Each compound was incubated with isolated human bone marrow cells cultured in semisolid medium over a period of 2 weeks. Venetoclax replicated the dose-dependent reduction in granulocyte colony formation caused by navitoclax, but

Cell type Cell line
Navitoclax EC50 (mM)
Venetoclax EC50 (mM)
A-1155463 EC50 (mM)
the BCL-XL–selective inhibitor A-1155463 had little or no effect (Fig. 6A). Likewise, navitoclax and venetoclax showed greater inhibition of gran- ulocyte colony formation than did A-1155463 when combined with

SCLC

 

 

 

 

 

 

AML
NCI-H889 NCI-H211 NCI-H187 NCI-H345 NCI-H446 NCI-H1048
NCI-H69 NCI-H847 NCI-H1417
DMS79 NCI-H1836
GDM-1
EOL-1 MOLM-13
HL-60 MV4-11
ML-2 SIG-M5
OCI-AML2 MOLM-16 OCI-AML5
THP-1 Kasumi-1
KG-1 HNT-34
PL-21
SKM-1
UKE-1
SET-2
HEL OCI-M2 OCI-M1
OCI-AML3 NOMO-1
ME-1
0.053
0.100
0.037
0.300
0.850
0.900
0.180
0.140
0.100
0.440
0.260
0.011
0.011
0.030
0.040
0.040
0.050
0.011
0.011
0.120
0.180
0.450
0.030
0.110
0.030
0.190
0.240
0.050
0.060
0.070
0.160
0.410
>5.000
>5.000
>5.000
0.015
0.330
0.850
1.200
2.400
2.700
3.860
3.900
4.100
4.620
5.000
0.100
0.100
0.100
0.100
0.100
0.100
0.100
0.100
0.040
0.110
0.820
1.000
1.200
1.700
1.810
2.530
3.060
3.330
3.370
4.050
>5.000
>5.000
>5.000
>5.000
6.310
1.350
0.180
>10.000 0.200
>10.000 3.200 0.003 0.007 1.440 0.015 1.170 2.250
>5.000
>5.000
>5.000
>5.000
>5.000
>5.000
>5.000
>5.000
>5.000 0.011
>5.000 0.680
>5.000
>5.000 0.490 0.080 0.120 0.100 0.230
>5.000
>5.000
>5.000
docetaxel (Fig. 6B).
To examine these effects in vivo, we evaluated orally bioavail- able BCL-2 family inhibitors, including A-874009 (a close analog of navitoclax), A-1211212 (a close analog of venetoclax), and A-1331852 (Fig. 4A). Male Sprague-Dawley rats were dosed daily with each com- pound for 5 days, either alone or in combination with a single dose of docetaxel (5 mg/kg), before assessing their circulating platelet and neutrophil counts. Docetaxel decreased neutrophil counts significantly, with reductions between 61% (P < 0.01) and 71% (P < 0.01) relative to the mean of the vehicle control groups (Fig. 6C). Rats treated with the dual BCL-2/BCL-XL inhibitor A-874009 or the BCL-2–selective in- hibitor A-1211212 had reduced neutrophils as well, in both cases showing a reduction of 41% relative to the respective vehicle-treated groups (P < 0.01). However, rats dosed with the BCL-XL–selective in- hibitor A-1331852 showed no inhibition of granulopoiesis and ex- hibited increased neutrophil counts. When combinations of docetaxel with A-874009, A-1211212, or A-1331852 were compared to docetaxel alone, the differences in neutrophil inhibition did not reach statis- tical significance (P = 0.383, 0.981, and 0.773, respectively). The dual BCL-2/BCL-XL and BCL-XL–selective inhibitors both induced significant reductions in circulating platelets (P < 0.01), though the BCL-2–selective compound A-1211212 did not (Fig. 6C). This is con- sistent with BCL-XL maintaining platelet survival and suggests that the lack of neutrophil inhibition was not due to insufficient expo- sure to A-1331852.
These studies indicate that BCL-2 inhibition accounts for the ex- acerbation of docetaxel-induced neutropenia caused by navitoclax (18) and support the hypothesis that BCL-XL–selective inhibitors will avoid dose-limiting neutropenia in this setting. However, because BCL-XL– selective inhibitors will still affect platelets, a question remained whether efficacious levels of BCL-XL inhibition can be achieved in combination with docetaxel before thrombocytopenia becomes dose-limiting. We therefore analyzed data from phase 1 clinical trials of navitoclax, in- cluding single-agent studies in subjects with lymphoid malignancies (11) or solid tumors (31) and a navitoclax-docetaxel combination study in subjects with advanced solid tumors (18). No apparent pharmaco- kinetic (PK) interactions were observed between navitoclax and docetaxel in the latter study (18). Moreover, coadministration with docetaxel did not lead to substantial increases in navitoclax-induced thrombocytopenia relative to navitoclax treatment alone (fig. S6). However, the maximum tolerated dose (MTD) of navitoclax when given in combination with docetaxel was limited to 150 mg/day because of febrile neutropenia, which was observed at navitoclax exposures as low as 50.7 mg*hour/ml (AUC0–inf) (18) (table S1). This is in contrast to the single-agent setting, where MTDs of 315 or 325 mg/day were achieved before thrombocy-

BCL-2 inhibition suppresses granulocyte colony formation and decreases circulating neutrophil counts
Although the navitoclax-docetaxel combination is efficacious in pre- clinical studies, neutropenia has limited the dosing of this combination
topenia became dose-limiting (11, 31). These doses correspond to mean exposures of 80.5 ± 44.3 mg*hour/ml and 91.0 ± 33.5 mg*hour/ml (AUC0–inf), which overlap with highly efficacious exposures of navitoclax determined in preclinical studies (7) (Table 4). Because the PK of
Table 3. Bliss synergy assessment of cell killing by BCL-2 family in- hibitors combined with docetaxel. Breast cancer, NSCLC, and ovarian cancer cell lines were cultured in the presence of navitoclax, A-1155463, or venetoclax plus or minus docetaxel for 72 hours before assessing cell viability. The sum of Bliss scores across the combination dose matrix is listed for each cell line.
Cell
type
Cell
line

H661

 

Navitoclax + docetaxel
109

Bliss sums A-1155463 +
docetaxel
94

 

Venetoclax +
docetaxel
-54

 

Cell
type

Breast cancer (9 × 3 dose matrix)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

NSCLC (5 × 5 dose
Cell
line

MDA-MB-231 MDA-MB-231 (LC3) MDA-MB-175 (VII)
HCC1806 MX-1
HCC1395 HCC1428
HCC38 ZR-75-1 HCC1569 HCC1500
MCF-7 (Luc) MDA-MB-361
SK-BR-3 UAC812 BT-474
HCC1187 HCC1937 DU4475 Hs578T T-47D
ZR-75-30 BT-20
HCC2218 HCC1590
BT-549 HCC1143
MDA-MB-436 LXFL529 SW1573

 

Navitoclax + docetaxel
264
253
-19 269
-43 80 153 67 271 193 165 100 66 83
-105 5 99 220 242 99 289
-290 126
-134 40 41 45
-110 656 354

Bliss sums A-1155463 +
docetaxel

277
305
290
224
53
-31
-203 42 103 185 146 63 110 101
-269
-5 0
332
229
62
250
-265 260
-250
-105 117 195
-96 734 461

 

Venetoclax +
docetaxel

-159
-62
-130
-46
-10
-109
-204
-201
-312
-102 15
-133
-40
-78
-210
-43
-173
-106 168
-161
-219
-418
-106
-135
-245
-76
-215
-386 236
0

 

 

 

Ovarian cancer (9 × 3 dose matrix)
H460
H1568
H838
EBC-1
H1703 OVCAR-3 OVCAR-4 OVCAR-5 OVCAR-8
SKOV3 IGROV1
128
-27 63
-107 1
635
-33 511 247 487 33
105
27
128
-103 29 683
-129 706 159 747
-49
15
-29
-6 18 6
-334
-231 92
-67
-53
-191

matrix)

 

 

 

 

continued
SKMES-1 A549 H1650 H1975 H1299
H23
H522
174
144
335
69
280
95
195
152
137
312
6
302
97
242
80
-19 146
-23 67 6
-19
Fig. 3. BCL-XL inhibition drives synergistic killing of solid tumor cell lines in combination with docetaxel. (A and B) Panels of (A) breast cancer (n = 28) and (B) NSCLC (n = 15) cell lines were cultured in the pres- ence of navitoclax, venetoclax, or A-1155463 plus or minus docetaxel for 72 hours before assessing cell viability. The sum of Bliss scores across the combination dose matrix was calculated for each cell line. Bliss sums were plotted for navitoclax-docetaxel combinations versus the venetoclax-docetaxel combinations or the A-1155463–docetaxel combinations. A significant correla- tion was observed between navitoclax-docetaxel and A-1155463–docetaxel Bliss sums for both the breast cancer (Spearman: 0.75, P < 0.0001) and NSCLC (Spearman: 0.94, P < 0.0001) cell line panels.
Fig. 4. Orally bioavailable inhibitor A-1331852 enables the functional character- ization of BCL-XL in vivo. (A) The chemical structures and selectivity profiles of BH3 mi- metics used for in vivo studies are depicted. A-874009 inhibits both BCL-2 and BCL-XL , whereas A-1211212 selectively inhibits BCL-2 and A-1331852 selectively inhibits BCL-XL. (B) Quantitative measurement of BCL-XL–BIM and BCL-2–BIM complexes in Molt-4 cells after 4-hour treatments with increasing concentrations of A-1331852. Data represent the average of triplicate experiments, with error bars indicating the SD. (C) Cytochrome c levels present in Molt-4 mitochondrial and cytosolic fractions as determined by immu- noblotting after 4-hour treatments with increasing concentrations of A-1331852. (D) Caspase-3/7 activation in Molt-4 cells af- ter incubation with increasing concentrations of A-1331852 for 4 hours. Data represent the average of triplicate experiments, with error bars indicating the SD. (E) Exposure of phos- phatidylserine as determined by annexin V staining of Molt-4 cells after 1-hour preincu- bation plus or minus the caspase inhibitor Z-VAD-fmk (75 mM) and an additional 24 hours of incubation with increasing concentra- tions of A-1331852. Data represent the av- erage of triplicate experiments, with error bars indicating the SD. (F) Mice bearing Molt-4 T cell acute lymphocytic leukemia– xenografted tumors were treated with a ve- hicle control (solid gray circles), venetoclax (daily for 14 days) at 100 mg/kg (open blue circles), or A-1331852 (twice a day for 14 days) at 25 mg/kg (open red circles). The points of each curve reflect the average volume of 10 tumors. Error bars indicate the SD of the means. (G) Mice bearing subcutaneous xe- nografts of NCI-H1963.FP5 were treated with vehicle control (solid gray circles), navitoclax (daily for 14 days) at 100 mg/kg (open green circles), venetoclax (daily for 14 days) at 50 mg/kg (open blue circles), A-1331852 (twice a day for 14 days) at 25 mg/kg (open red circles), or a combination of the latter two compounds (solid purple circles). The points of each curve reflect the average volume of five tumors. Error bars indicate the SD.

 

 

 

 

 
navitoclax is linear over the 150- to 325-mg dose range (11), these data indicate that large increases in navitoclax exposure could be obtained in combination with docetaxel if neutropenia were avoided. The dis-
covery of BCL-XL–selective inhibitors like A-1331852 thus represents an opportunity to maximize BCL-XL inhibition and improve efficacy when dosed in combination with docetaxel.
Table 4. Inhibition of tumor xenograft growth by administration of A-1331852, docetaxel, or the combination. TGImax = 100 (1 – Tv/Cv), where Tv and Cv are the mean tumor volumes of the treated and control groups, respectively. TGD is the extended period of time that a treated tumor re- quires to reach a volume of 1 cm3 relative to the control group. TGD = 100

(T/C – 1), where T and C are the median time periods required for the treated and control groups, respectively, to reach 1 cm3. Each treatment group of NCI-H747, A549-FP3, EBC-1, and OVCAR-5 models consisted of five mice. Each treatment group of the NCI-H358 model consisted of eight mice. Each treatment group of NCI-H1650 and MDA-MB-231 LC3 consisted of ten mice.
A-1331852 Docetaxel A-1331852 + docetaxel
Xenograft
TGImax TGD TGImax TGD TGImax TGD
NCI-H1650 48* 33* 82* 52* 98*† 170*†
MDA-MB-231 LC3 55* 55* 73* 82* 93*† 141*†
NCI-H747 50* 33* 52* 113* 84* 167*†
A549-FP3 67* 108* 71* 108* 74* 131*
EBC-1 52* 22* 76* 44* 88*† 94*†
OVCAR-5 34* 54* 39* 54 57*† 85*†
NCI-H358 62* 80* 58* 47* 91*† 233*†

*Significantly different from sham-treated control group. †Significantly different from most effective component of the combination as a single agent.

 

ABT-737 and navitoclax (6, 7), came the ability to inhibit BCL-2 and BCL-XL di- rectly with cell-permeable small molecules, facilitating a host of new discoveries. Since its introduction, the BCL-2–selective in- hibitor ABT-199/GDC-0199 (venetoclax) has been used to define the role of BCL-2 in specific hematopoietic lineages (32), a va- riety of hematologic cancers (25, 33–38), and even some solid tumors, including es- trogen receptor–positive breast cancer (39). In addition, certain roles for BCL-XL have been deduced using a subtractive parsing method that compares the effects of ABT- 737 or navitoclax to those of venetoclax (38, 40–42). For example, Bah and co-workers used ABT-737 and venetoclax to demon- strate that BCL-XL inhibition is required

Fig. 5. BCL-XL–selective inhibitor A-1331852 enhances the efficacy of docetaxel in vivo. The growth inhibition of established tumors in SCID-bg mice is illustrated. Each graph describes the change of tumor volume (mm3, ordinate) as a function of the time after initiation of treatment (days, abscissa). Each point on the curves represents the mean volume of 10 tumors. Error bars depict the SD. A-1331852 was administered orally and docetaxel was administered intravenously for all studies. (A) MDA-MB-231 LC3 metastatic breast cancer xenograft. Vehicle control (solid gray circles), docetaxel (DTX) (once) at 7.5 mg/kg (open blue circles), A-1331852 (daily for 14 days) at 25 mg/kg (open red circles), and combination of docetaxel (once) at 7.5 mg/kg and A-1331852 (daily for 14 days) at 25 mg/kg (solid purple circles). (B) NCI-H1650 NSCLC xenograft. Vehicle control (solid gray circles), docetaxel (once) at 7.5 mg/kg (open blue circles), A-1331852 (daily for 14 days) at 25 mg/kg (open red circles), and combination of docetaxel (once) at 7.5 mg/kg and A-1331852 (daily for 14 days) at 25 mg/kg (solid purple circles).
for the killing of breast cancer cell lines in combination with paclitaxel (41). Although this approach is informative, these inhibi- tors alone cannot distinguish between cell lines that are co-dependent on BCL-2 and BCL-XL for survival and those that rely solely on BCL-XL for survival. With the generation of the BCL-XL–selective inhib- itors described here, it is possible to inter- rogate the role of BCL-XL directly in vitro and in vivo and to determine whether its selective inhibition is sufficient for a giv-

DISCUSSION

The work that led to the discovery and functional characterization of BCL-2 family proteins relied on a variety of cell and molecular biology approaches, genetically engineered mouse models, and RNA interference– based methods. With the synthesis of the first validated BH3 mimetics,
en effect. This is demonstrated by the studies in Fig. 2, which define the contributions of BCL-2 and BCL-XL in maintaining the survival of SCLC and AML cell lines. When used together, the small-molecule BH3 mimetics described here represent a powerful toolkit for dissect- ing the roles of BCL-2 and BCL-XL and for defining more effective therapeutic strategies. As these molecules find broader use in the research
Fig. 6. BCL-2–selective inhibition suppresses granulopoiesis ex vivo and reduces circulating neutrophil counts in vivo. (A) Isolated human bone marrow cells were used to perform granulocyte colony-forming as- says in the presence of increasing concentrations of navitoclax (BCL-2/BCL-XL), venetoclax (BCL-2–selective), or A-1155463 (BCL-XL–selective). Neutrophil colonies were enumerated by light microscopy after 15 to 16 days of cul- ture. (B) Neutrophil colony formation was assessed as in (A) after treatment of human bone marrow samples with increasing concentrations of navitoclax, venetoclax, or A-1155463 ± 5 nM docetaxel. Colonies were enumerated after 14 days of growth in culture. All data in (A) and (B) represent the means of triplicate experiments, with error bars indicating the SD. Student’s t tests compared test compound single-agent effects to solvent control sam- ples. For combination experiments, test compounds in combination with docetaxel were compared to samples treated with docetaxel alone. (C) Groups (n = 10 per group) of male Sprague-Dawley rats were dosed with docetaxel (5 mg/kg, intravenously, once), the BCL-2/BCL-XL inhibitor A-874009 (30 mg/kg, orally, daily for 5 days), the BCL-2–selective inhibitor A-1211212 (50 mg/kg, orally, daily for 5 days), the BCL-XL–selective inhibitor A-1331852 (7 mg/kg, oral- ly, daily for 5 days), or their respective vehicles. Groups were also dosed with combinations of docetaxel and the various BCL-2 family inhibitors. Box and whisker plots depict the means (n = 10) and ranges for total neu- trophil and platelet counts assessed in blood collected on day 6. Tukey- Kramer tests compared vehicle- and compound-treated groups at a 5% significance level. Significant reductions in neutrophils or platelets relative to vehicle controls (P < 0.01) are denoted by asterisks.

 

community, they should facilitate a more complete understanding of BCL-2 family biology and the roles these proteins play in normal tissues and disease states.
Antiapoptotic BCL-2 family proteins have been widely studied in cancer cells, where they are often overexpressed and maintain survival by sequestering large amounts of their proapoptotic counterparts, a state that has been referred to as being primed for death (5). An ele- gant peptide-based technique referred to as “BH3 profiling” has been used extensively to probe these interactions and to determine the pro- file of BCL-2 family dependence for a host of tumor cell types (43). One limitation of this method is the lack of highly selective BH3 pep- tides for targeting certain BCL-2 family members, which can make it difficult to fully define their roles. The discovery of selective BH3 mi- metics like venetoclax, A-1155463, A-1331852, and cell-active MCL- 1–selective inhibitors (44) will now enable a form of chemical BH3 profiling, whereby one can determine precisely which BCL-2 family members are required for the survival of any given cell population. The ability of these molecules to penetrate live cells also obviates the need for detergent-based permeabilization associated with peptide- based methods and enables one to probe BCL-2 family dependency by conducting simple cell viability assays. Furthermore, navitoclax, venetoclax, and A-1331852 are orally bioavailable and sufficiently po- tent to interrogate BCL-2 family member dependence in vivo, making them especially attractive translational tools.
Despite major investments into the discovery and development of small-molecule therapeutics, many compounds still fail in the clinic be- cause of a lack of efficacy (45). In some cases, this may be due to off-target activities causing dose-limiting toxicities that preclude the

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 
solid tumors. BCL-XL–selective inhibition enhanced the efficacy of docetaxel in breast cancer, NSCLC, and ovarian cancer models, but it did not inhibit granulopoiesis assessed ex vivo or in vivo. In con- trast, BCL-2–selective inhibition suppressed granulopoiesis to a similar extent as was observed with dual BCL-2/BCL-XL inhibitors. Although Bcl-2 knockout mice show no overt signs of defective granulopoiesis before succumbing to polycystic kidney disease (46, 47), bone marrow

achievement of efficacious exposures. This emphasizes the need to
-/-
from Bcl-2
reconstituted mice shows a reduced capacity for granu-

carefully dissect a molecule’s target inhibition profile and to better un- derstand the roles each target plays in efficacy and toxicity. As demon- strated here, selective BCL-XL inhibitors can offer important advantages over their less selective predecessor, navitoclax, for the treatment of
locyte colony formation, especially when cultured in the absence of cytokines such as interleukin-3 and stem cell factor (48). Moreover, neutropenia has been observed in a portion of CLL subjects receiving venetoclax (16).
Navitoclax clinical data indicated that, if exacerbated neutropenia in combination with docetaxel could be avoided, higher exposures and more complete inhibition of BCL-XL might be achieved before throm- bocytopenia becomes dose-limiting. However, these analyses were limited by the small number of subjects (n = 7) with both exposure and hematology data available. In addition, our in vivo studies of BCL-XL– selective inhibitors were limited to rodent models, and so the possibil- ity of encountering dose-limiting toxicities other than thrombocytopenia in the clinic cannot be ruled out. Nevertheless, the data presented here indicate that BCL-XL–selective inhibitors have the potential for im- proved safety and efficacy profiles, and provide further impetus for ex- ploring this concept in the clinic.

SUPPLEMENTARY MATERIALS

www.sciencetranslationalmedicine.org/cgi/content/full/7/279/279ra40/DC1 Materials and Methods

itor of BCL-2, in lymphoid malignancies: A phase 1 dose-escalation study of safety, pharma- cokinetics, pharmacodynamics, and antitumour activity. Lancet Oncol. 11, 1149–1159 (2010).
12.A. W. Roberts, J. F. Seymour, J. R. Brown, W. G. Wierda, T. J. Kipps, S. L. Khaw, D. A. Carney, S. Z. He, D. C. S. Huang, H. Xiong, Y. Cui, T. A. Busman, E. M. McKeegan, A. P. Krivoshik, S. H. Enschede, R. Humerickhouse, Substantial susceptibility of chronic lymphocytic leukemia to BCL2 inhibi- tion: Results of a phase I study of navitoclax in patients with relapsed or refractory disease. J. Clin. Oncol. 30, 488–496 (2012).
13.H. Zhang, P. M. Nimmer, S. K. Tahir, J. Chen, R. M. Fryer, K. R. Hahn, L. A. Iciek, S. J. Morgan, M. C. Nasarre, R. Nelson, L. C. Preusser, G. A. Reinhart, M. L. Smith, S. H. Rosenberg, S. W. Elmore, C. Tse, Bcl-2 family proteins are essential for platelet survival. Cell Death Differ. 14, 943–951 (2007).
14.K. D. Mason, M. R. Carpinelli, J. I. Fletcher, J. E. Collinge, A. A. Hilton, S. Ellis, P. N. Kelly, P. G. Ekert, D. Metcalf, A. W. Roberts, D. C. S. Huang, B. T. Kyle, Programmed anuclear cell death delimits platelet life span. Cell 128, 1173–1186 (2007).
15.A. J. Souers, J. D. Leverson, E. R. Boghaert, S. L. Ackler, N. D. Catron, J. Chen, B. D. Dayton, H. Ding, S. H. Enschede, W. J. Fairbrother, D. C. S. Huang, S. G. Hymowitz, S. Jin, S. L. Khaw, P. J. Kovar, L. T. Lam, J. Lee, H. L. Maecker, K. C. Marsh, K. D. Mason, M. J. Mitten, P. M. Nimmer, A. Oleksijew, C. H. Park, C. M. Park, D. C. Phillips, A. W. Roberts, D. Sampath, J. F. Seymour, M. L. Smith, G. M. Sullivan, S. K. Tahir, C. Tse, M. D. Wendt, Y. Xiao, J. C. Xue, H. Zhang, R. A. Humerickhouse, S. H. Rosenberg, S. W. Elmore, ABT-199, a potent and selective BCL-2 inhibitor, achieves anti- tumor activity while sparing platelets. Nat. Med. 19, 202–208 (2013).
16.J. F. Seymour, M. S. Davids, J. M. Pagel, B. S. Kahl, W. G. Wierda, S. Puvvada, J. F. Gerecitano,

-/-
Fig. S1. BCL-XL–selective inhibitor A-1155463 kills Mcl-1
-/-
MEF cells but not Bak
-/-
Bax
MEFs.
T. J. Kipps, M. A. Anderson, D. C. S. Huang, N. Rudersdorf, L. A. Gressick, N. P. Montalvo, J. Yang,

Fig. S2. Navitoclax synergizes with docetaxel to kill breast cancer cell lines.
Fig. S3. Navitoclax-docetaxel combination kills breast cancer cell lines by inducing apoptosis. Fig. S4. Selective BCL-XL inhibition suffices for synergy with docetaxel in ovarian cancer cell lines.
M. Zhu, M. Dunbar, E. Cerri, S. H. Enschede, R. Humerickhouse, A. W. Roberts, ABT-199 (GDC-0199) in relapsed/refractory (R/R) chronic lymphocytic leukemia (CLL) and small lymphocytic lymphoma (SLL): High complete response rate and durable disease control. J. Clin. Oncol. 32, 448s (2014).

-/-
Fig. S5. BCL-XL–selective inhibitor A-1331852 kills Mcl-1
-/-
MEF cells but not Bak
-/-
Bax
MEFs.
17.M. S. Davids, J. F. Seymour, J. F. Gerecitano, B. S. Kahl, J. M. Pagel, W. G. Wierda, M. A. Anderson,

Fig. S6. Relationship between exposure and platelet reduction is similar for navitoclax with or without docetaxel.
Table S1. Plasma exposures and platelet effects of navitoclax-docetaxel combinations. References (49–51)
REFERENCES AND NOTES

1.D. Hanahan, R. A. Weinberg, Hallmarks of cancer: The next generation. Cell 144, 646–674 (2011).
2.J. E. Chipuk, T. Moldoveanu, F. Llambi, M. J. Parsons, D. R. Green, The BCL-2 family reunion. Mol. Cell 37, 299–310 (2010).
3.X. Chi, J. Kale, B. Leber, D. W. Andrews, Regulating cell death at, on, and in membranes. Biochim. Biophys. Acta, 1843, 2100–2113 (2014).
4.P. E. Czabotar, G. Lessene, A. Strasser, J. M. Adams, Control of apoptosis by the BCL-2 protein family: Implications for physiology and therapy. Nat. Rev. Mol. Cell Biol. 15, 49–63 (2014).
5.M. Certo, V. Del Gaizo Moore, M. Nishino, G. Wei, S. Korsmeyer, S. A. Armstrong, A. Letai, Mitochondria primed by death signals determine cellular addiction to antiapoptotic BCL-2 family members. Cancer Cell 9, 351–365 (2006).
6.T. Oltersdorf, S. W. Elmore, A. R. Shoemaker, R. C. Armstrong, D. J. Augeri, B. A. Belli, M. Bruncko, T. L. Deckwerth, J. Dinges, P. J. Hajduk, M. K. Joseph, S. Kitada, S. J. Korsmeyer, A. R. Kunzer, A. Letai, C. Li, M. J. Mitten, D. G. Nettesheim, S. Ng, P. M. Nimmer, J. M. O’Connor, A. Oleksijew, A. M. Petros, J. C. Reed, W. Shen, S. K. Tahir, C. B. Thompson, K. J. Tomaselli, B. Wang, M. D. Wendt, H. Zhang, S. W. Fesik, S. H. Rosenberg, An inhibitor of Bcl-2 family proteins induces regression of solid tumours. Nature 435, 677–681 (2005).
7.C. Tse, A. R. Shoemaker, J. Adickes, M. G. Anderson, J. Chen, S. Jin, E. F. Johnson, K. C. Marsh, M. J. Mitten, P. Nimmer, L. Roberts, S. K. Tahir, Y. Xiao, X. Yang, H. Zhang, S. Fesik, S. H. Rosenberg, S. W. Elmore, ABT-263: A potent and orally bioavailable Bcl-2 family inhibitor. Cancer Res. 68, 3421–3428 (2008).
8.A. R. Shoemaker, M. J. Mitten, J. Adickes, S. Ackler, M. Refici, D. Ferguson, A. Oleksijew, M. O’Connor, B. Wang, D. J. Frost, J. Bauch, K. Marsh, S. K. Tahir, X. Yang, C. Tse, S. W. Fesik, S. H. Rosenberg, S. W. Elmore, Activity of the Bcl-2 family inhibitor ABT-263 in a panel of small cell lung cancer xenograft models. Clin. Cancer Res. 14, 3268–3277 (2008).
9.S. Ackler, M. J. Mitten, K. Foster, A. Oleksijew, M. Refici, S. K. Tahir, Y. Xiao, C. Tse, D. J. Frost, S. W. Fesik, S. H. Rosenberg, S. W. Elmore, and A. R. Shoemaker. The Bcl-2 inhibitor ABT-263 enhances the response of multiple chemotherapeutic regimens in hematologic tumors in vivo. Cancer Chemother. Pharmacol. 66, 869–880 (2010).
10.J. Chen, S. Jin, V. Abraham, X. Huang, B. Liu, M. Mitten, P. Nimmer, X. Lin, M. Smith, Y. Shen, A. R. Shoemaker, S. K. Tahir, H. Zhang, S. L. Ackler, S. H. Rosenberg, H. Maecker, D. Sampath, J. D. Leverson, C. Tse, S. W. Elmore, The Bcl-2/Bcl-XL/Bcl-w inhibitor, navitoclax, enhances the activity of chemotherapeutic agents in vitro and in vivo. Mol. Cancer Ther. 10, 2340–2349 (2011).
11.W. H. Wilson, O. A. O’Connor, M. S. Czuczman, A. S. LaCasce, J. F. Gerecitano, J. P. Leonard, A. Tulpule, K. Dunleavy, H. Xiong, Y.-L. Chiu, Y. Cui, T. Busman, S. W. Elmore, S. H. Rosenberg, A. P. Krivoshik, S. H. Enschede, R. A. Humerickhouse, Navitoclax, a targeted high-affinity inhib-
N. Rudersdorf, L. A. Gressick, N. P. Montalvo, J. Yang, M. Zhu, M. Dunbar, E. Cerri, S. H. Enschede, R. Humerickhouse, A. W. Roberts, Phase I study of ABT-199 (GDC-0199) in patients with relapsed/
refractory (R/R) non-Hodgkin lymphoma (NHL): Responses observed in diffuse large B-cell (DLBCL) and follicular lymphoma (FL) at higher cohort doses. J. Clin. Oncol. 32, 544s (2014).
18.M. Puglisi, L. van Doorn, M. Blanco-Codesido, M. J. De Jonge, K. Moran, J. Yang, T. Busman, C. Franklin, M. Mabry, A. Krivoshik, R. Humerickhouse, L. R. Molife, F. Eskens, A phase I safety and pharmacokinetic (PK) study of navitoclax (N) in combination with docetaxel (D) in patients (pts) with solid tumors. J. Clin. Oncol. 29, 2518 (2011).
19.Z.-F. Tao, L. Hasvold, L. Wang, X. Wang, A. M. Petros, C. H. Park, E. R. Boghaert, N. D. Catron, J. Chen, P. M. Colman, P. E. Czabotar, K. Deshayes, W. J. Fairbrother, J. A. Flygare, S. G. Hymowitz, S. Jin, R. A. Judge, M. F. Koehler, P. J. Kovar, G. Lessene, M. J. Mitten, C. O. Ndubaku, P. Nimmer, H. E. Purkey, A. Oleksijew, D. C. Phillips, B. E. Sleebs, B. J. Smith, M. L. Smith, S. K. Tahir, K. G. Watson, Y. Xiao, J. Xue, H. Zhang, K. Zobel, S. H. Rosenberg, C. Tse, J. D. Leverson, S. W. Elmore, A. J. Souers, Discovery of a potent and selective BCL-XL inhibitor with in vivo activity. ACS Med. Chem. Lett. 5, 1088–1093 (2014).
20.G. Lessene, P. E. Czabotar, B. E. Sleebs, K. Zobel, K. N. Lowes, J. M. Adams, J. B. Baell, P. M. Colman, K. Deshayes, W. J. Fairbrother, J. A. Flygare, P. Gibbons, W. J. A. Kirsten, S. Kulasegaram, R. M. Moss, J. B. Parisot, B. J. Smith, I. P. Street, H. Yang, D. C. S. Huang, K. G. Watson, Structure-guided design of a selective BCL-XL inhibitor. Nat. Chem. Biol. 9, 390–397 (2013).
21.S. N. Willis, L. Chen, G. Dewson, A. Wei, E. Naik, J. I. Fletcher, J. M. Adams, D. C. S. Huang, Proapoptotic Bak is sequestered by Mcl-1 and Bcl-xL, but not Bcl-2, until displaced by BH3-only proteins. Genes Dev. 19, 1294–1305 (2005).
22.M. F. van Delft, A. H. Wei, K. D. Mason, C. J. Vandenberg, L. Chen, P. E. Czabotar, S. N. Willis, C. L. Scott, C. L. Day, S. Cory, J. M. Adams, A. W. Roberts, D. C. S. Huang, The BH3 mimetic ABT-737 targets selective Bcl-2 proteins and efficiently induces apoptosis via Bak/Bax if Mcl-1 is neutralized. Cancer Cell 10, 389–399 (2006).
23.S. K. Tahir, X. Yang, M. G. Anderson, S. E. Morgan-Lappe, A. V. Sarthy, J. Chen, R. B. Warner, S.-C. Ng, S. W. Fesik, S. H. Rosenberg, S. W. Elmore, C. Tse, Influence of Bcl-2 family mem- bers on the cellular response of small-cell lung cancer cell lines to ABT-737. Cancer Res. 67, 1176–1183 (2007).
24.E. T. Olejniczak, C. Van Sant, M. G. Anderson, G. Wang, S. K. Tahir, G. Sauter, R. Lesniewski, D. Semizarov, Integrative genomic analysis of small-cell lung carcinoma reveals correlates of sensitivity to Bcl-2 antagonists and uncovers novel chromosomal gains. Mol. Cancer Res. 5, 331–339 (2007).
25.R. Pan, L. J. Hogdal, J. M. Benito, D. Bucci, L. Han, G. Borthakur, J. Cortes, D. J. DeAngelo, L. Debose, H. Mu, H. Döhner, V. I. Gaidzik, I. Galinsky, L. S. Golfman, T. Haferlach, K. G. Harutyunyan, J. Hu, J. D. Leverson, G. Marcucci, M. Müschen, R. Newman, E. Park, P. P. Ruvolo, V. Ruvolo, J. Ryan, S. Schindela, P. Zweidler-McKay, R. M. Stone, H. Kantarjian, M. Andreeff, M. Konopleva, A. G. Letai, Selective BCL-2 inhibition by ABT-199 causes on target cell death in acute myeloid leukemia. Cancer Discov. 4, 362–375 (2014).
26.N. Tan, M. Malek, J. Zha, P. Yue, R. Kassees, L. Berry, W. J. Fairbrother, D. Sampath, L. D. Belmont, Navitoclax enhances the efficacy of taxanes in non–small cell lung cancer models. Clin. Cancer Res. 17, 1394–1404 (2011).
27.S. R. Oakes, F. Vaillant, E. Lim, L. Lee, K. Breslin, F. Feleppa, S. Deb, M. E. Ritchie, E. Takano, T. Ward, S. B. Fox, D. Generali, G. K. Smyth, A. Strasser, D. C. S. Huang, J. E. Visvader, G. J. Lindeman,
Sensitization of BCL-2–expressing breast tumors to chemotherapy by the BH3 mimetic ABT-737. Proc. Natl. Acad. Sci. U.S.A. 109, 2766–2771 (2012).
28.M. Wong, N. Tan, J. Zha, F. V. Peale, P. Yue, W. J. Fairbrother, L. D. Belmont, Navitoclax (ABT-263) reduces Bcl-xL–mediated chemoresistance in ovarian cancer models. Mol. Cancer Ther. 11, 1026–1035 (2012).
29.M. C. Berenbaum, Criteria for analyzing interactions between biologically active agents. Adv. Cancer Res. 35, 269–335 (1981).
30.A. A. Borisy, P. J. Elliott, N. W. Hurst, M. S. Lee, J. Lehár, E. R. Price, G. Serbedzija, G. R. Zimmermann, M. A. Foley, B. R. Stockwell, C. T. Keith, Systematic discovery of multicomponent therapeutics. Proc. Natl. Acad. Sci. U.S.A. 100, 7977–7982 (2003).
31.L. Gandhi, D. R. Camidge, M. Ribeiro de Oliveira, P. Bonomi, D. Gandara, D. Khaira, C. L. Hann, E. M. McKeegan, E. Litvinovich, P. M. Hemken, C. Dive, S. H. Enschede, C. Nolan, Y.-L. Chiu, T. Busman, H. Xiong, A. P. Krivoshik, R. Humerickhouse, G. I. Shapiro, C. M. Rudin, Phase I study of navitoclax (ABT-263), a novel Bcl-2 family inhibitor, in patients with small-cell lung cancer and other solid tumors. J. Clin. Oncol. 29, 909–916 (2011).
32.S. L. Khaw, D. Mérino, M. A. Anderson, S. P. Glaser, P. Bouillet, A. W. Roberts, D. C. S. Huang, Both leukaemic and normal peripheral B lymphoid cells are highly sensitive to the selective pharmacological inhibition of prosurvival Bcl-2 with ABT-199. Leukemia 28, 1207–1215 (2014).
33.C. J. Vandenberg, S. Cory, ABT-199, a new Bcl-2–specific BH3 mimetic, has in vivo efficacy against aggressive Myc-driven mouse lymphomas without provoking thrombocytopenia. Blood 121, 2285–2288 (2013).
34.C. Touzeau, C. Dousset, S. Le Gouill, D. Sampath, J. D. Leverson, A. J. Souers, S. Maïga, M. C. Béné, P. Moreau, C. Pellat-Deceunynck, M. Amiot, The Bcl-2 specific BH3 mimetic ABT-199: A promising targeted therapy for t(11;14) multiple myeloma. Leukemia 28, 210–212 (2014).
35.X. Niu, G. Wang, Y. Wang, J. T. Caldwell, H. Edwards, C. Xie, J. W. Taub, C. Li, H. Lin, Y. Ge, Acute myeloid leukemia cells harboring MLL fusion genes or with the acute promyelocytic leukemia phenotype are sensitive to the BCL-2-selective inhibitor ABT-199. Leukemia 28, 1557–1560 (2014).
36.J. M. Bogenberger, D. Delman, N. Hansen, R. Valdez, V. Fauble, R. A. Mesa, R. Tibes, Ex vivo activity of BCL-2 family inhibitors ABT-199 and ABT-737 combined with 5-azacytidine in myeloid malignancies. Leuk. Lymphoma 56, 226–229 (2015).
37.N. M. Anderson, I. Harrold, M. R. Mansour, T. Sanda, M. McKeown, N. Nagykary, J. E. Bradner, G. Lan Zhang, A. T. Look, H. Feng, BCL2-specific inhibitor ABT-199 synergizes strongly with cytarabine against the early immature LOUCY cell line but not more differentiated T-ALL cell lines. Leukemia 28, 1145–1148 (2014).
38.T. N. Chonghaile, J. E. Roderick, C. Glenfield, J. Ryan, S. E. Sallan, L. B. Silverman, M. L. Loh, S. P. Hunger, B. Wood, D. J. DeAngelo, R. Stone, M. Harris, A. Gutierrez, M. A. Kelliher, A. Letai, Maturation stage of T-cell acute lymphoblastic leukemia determines BCL-2 versus BCL-XL dependence and sensitivity to ABT-199. Cancer Discov. 4, 1074–1087 (2014).
39.F. Vaillant, D. Merino, L. Lee, K. Breslin, B. Pal, M. E. Ritchie, G. K. Smyth, M. Christie, L. J. Phillipson, C. J. Burns, G. B. Mann, J. E. Visvader, G. J. Lindeman, Targeting BCL-2 with the BH3 mimetic ABT-199 in estrogen receptor-positive breast cancer. Cancer Cell 24, 120–129 (2013).
40.A. Basu, N. E. Bodycombe, J. H. Cheah,, E. V. Price, K. Liu, G. I. Schaefer, R. Y. Ebright, M. L. Stewart, D. Ito, S. Wang, A. L. Bracha, T. Liefeld, M. Wawer, J. C. Gilbert, A. J. Wilson, N. Stransky, G. V. Kryukov, V. Dancik, J. Barretina, L. A. Garraway, C. S.-Y. Hon, B. Munoz, J. A. Bittker, B. R. Stockwell, D. Khabele, A. M. Stern, P. A. Clemons, A. F. Shamji, S. L. Schreiber, An interactive resource to identify cancer genetic and lineage dependencies targeted by small molecules. Cell 154, 1151–1161 (2013).
41.N. Bah, L. Maillet, J. Ryan, S. Dubreil, F. Gautier, A. Letai, P. Juin, S. Barillé-Nion, Bcl-xL controls a switch between cell death modes during mitotic arrest. Cell Death Dis. 5, e1291 (2014).
42.A. Zeuner, F. Francescangeli, P. Contavalli, G. Zapparelli, T. Apuzzo, A. Eramo, M. Baiocchi, M. L. De Angelis, M. Biffoni, G. Sette, M. Todaro, G. Stassi, R. De Maria, Elimination of quiescent/
slow-proliferating cancer stem cells by Bcl-XL inhibition in non-small cell lung cancer. Cell Death Differ. 21, 1877–1888 (2014).
43.V. Del Gaizo Moore, A. Letai, BH3 profiling—Measuring integrated function of the mito- chondrial apoptotic pathway to predict cell fate decisions. Cancer Lett. 332, 202–205 (2013).
44.J. D. Leverson, H. Zhang, J. Chen, S. K. Tahir, D. C. Phillips, J. Xue, P. Nimmer, S. Jin, M. Smith, Y. Xiao, P. Kovar, A. Tanaka, M. Bruncko, G. S. Sheppard, L. Wang, S. Gierke, L. Kategaya, D. J. Anderson, C. Wong, J. Eastham-Anderson, M. J. C. Ludlam, D. Sampath, W. J. Fairbrother, I. Wertz, S. H. Rosenberg, C. Tse, S. W. Elmore, A. J. Souers, Potent and selective small-molecule MCL-1 inhibitors dem- onstrate on-target cancer cell killing activity as single agents and in combination with ABT-263 (navitoclax). Cell Death Dis. 6, e1590 (2015).

45.M. Hay, D. W. Thomas, J. L. Craighead, C. Economides, J. Rosenthal, Clinical development success rates for investigational drugs. Nat. Biotechnol. 32, 40–51 (2014).
46.D. J. Veis, C. M. Sorenson, J. R. Shutter, S. J. Korsmeyer, Bcl-2-deficient mice demonstrate fulminant lymphoid apoptosis, polycystic kidneys, and hypopigmented hair. Cell 75, 229–240 (1993).
47.K. Nakayama, K. Nakayama, I. Negishi, K. Kuida, H. Sawa, D. Y. Loh, Targeted disruption of Bcl-2ab in mice: Occurrence of gray hair, polycystic kidney disease, and lymphocytopenia. Proc. Natl. Acad. Sci. U.S.A. 91, 3700–3704 (1994).
48.A. Villunger, C. Scott, P. Bouillet, A. Strasser, Essential role for the BH3-only protein Bim but redundant roles for Bax, Bcl-2, and Bcl-w in the control of granulocyte survival. Blood 101, 2393–2400 (2003).
49.L. S. Kim, S. Huang, W. Lu, D. Chelouche Lev, J. E. Price, Vascular endothelial growth factor expression promotes the growth of breast cancer brain metastases in nude mice. Clin. Exp. Metastasis 21, 107–118 (2004).
50.Z.-X. Wang, An exact mathematical expression for describing competitive binding of two different ligands to a protein molecule. FEBS Lett. 360, 111–114 (1995).
51.D. C. Phillips, S. P. Garrison, J. R. Jeffers, G. P. Zambetti, Assays to measure p53-dependent and -independent apoptosis. Methods Mol. Biol. 559, 143–159 (2009).

Acknowledgments: We thank J. Damen and M. Huber of StemCell Technologies for technical advice and performing granulocyte colony-forming assays. We thank D. Sheinson for performing statistical analyses. We also thank M. Mabry for helpful discussions and coining the phrase “chemical parsing.” Funding: Work in the Huang Lab is supported by grants and fellowships from the Australian National Health and Medical Research Council (research fellowships to D.C.S.H.; project grants to D.C.S.H.; program grants 461219, 461221, and 1016701; and Independent Research Institutes Infrastructure Support Scheme grant 361646), the Cancer Council Victoria (grant-in-aid to D.C.S.H.), the Leukemia and Lymphoma Society (Specialized Centers of Research grants), the Australian Cancer Research Foundation, and a Victorian State Government Operational Infrastructure Support grant. Author contribu- tions: J.D.L. designed experiments, oversaw biology experiments, and wrote the manuscript. D.C.P. designed and performed breast cancer experiments in vitro, oversaw biology experiments, and wrote the manuscript. M.J.M., A.O., and T.J.M. performed in vivo efficacy experiments. K.S.V., E.R.B., D.H.A., and D.S. designed, oversaw, and analyzed in vivo efficacy experiments. D.D. de- signed and oversaw rat studies. J.M.T. and N.L. performed rat studies. L.D.B. designed and oversaw NSCLC experiments in vitro. K.N.L. performed ovarian cancer studies in vitro. P.K. performed bio- chemical affinity assessments of all compounds. Y.X. performed breast cancer experiments in vitro. X.M.M. performed AML experiments in vitro. P.N., S.J., J.X., J.C., and H.Z. performed biological characteriza- tions of BCL-XL–selective inhibitors in vitro. M.S. and S.K.T. designed colony-forming experiments and performed SCLC experiments in vitro. L.W., Z.-F.T., and M.D.W. designed and synthesized compounds. C.T., D.C.S.H., and W.J.F. oversaw biology efforts. S.H.R. and S.W.E. oversaw chemistry and biology efforts. A.J.S. designed compounds, oversaw chemistry and biology efforts, and wrote the manuscript. All authors contributed to the interpretation of data and to the review and editing of the manuscript. Competing interests: J.D.L., D.C.P., M.J.M., E.R.B., J.C., P.N., S.K.T., J.X., P.K., M.S., A.O., T.J.M., K.S.V., D.H.A., Y.X., H.Z., L.W., Z.T., M.D.W., C.T., S.H.R., S.W.E., and A.J.S. are employees of AbbVie and hold company stock. D.D., L.D.B., J.M.T., N.L., D.S., and W.J.F. are employees of Genentech and hold company stock. Financial support for this research was provided by AbbVie and Genentech. AbbVie, Genentech, and Walter and Eliza Hall Institute of Medical Research participated in the design and conduct of studies, interpretation of data, and review and approval of the publication. Patents cover all of the small mo- lecules described in this study. Data and materials availability: ABT-263 (navitoclax), ABT-199 (ve- netoclax), A-874009, A-1155463, A-1211212, and A-1331852 may be obtained through appropriate material transfer agreements. Please address compound requests to [email protected].

Submitted 9 December 2014 Accepted 13 February 2015 Published 18 March 2015 10.1126/scitranslmed.aaa4642

Citation: J. D. Leverson, D. C. Phillips, M. J. Mitten, E. R. Boghaert, D. Diaz, S. K. Tahir, L. D. Belmont, P. Nimmer, Y. Xiao, X. M. Ma, K. N. Lowes, P. Kovar, J. Chen, S. Jin, M. Smith, J. Xue, H. Zhang, A. Oleksijew, T. J. Magoc, K. S. Vaidya, D. H. Albert, J. M. Tarrant, N. La, L. Wang, Z.-F. Tao, M. D. Wendt, D. Sampath, S. H. Rosenberg, C. Tse, D. C. S. Huang, W. J. Fairbrother, S. W. Elmore, A. J. Souers, Exploiting selective BCL-2 family inhibitors to dissect cell survival dependencies and define improved strategies for cancer therapy. Sci. Transl. Med. 7, 279ra40 (2015).

Exploiting selective BCL-2 family inhibitors to dissect cell survival dependencies and define improved strategies for cancer therapy Joel D. Leverson et al.
Sci Transl Med 7, 279ra40 (2015); DOI: 10.1126/scitranslmed.aaa4642

 

Editor’s Summary

A more refined antitumor strategy
The BCL-2 family is a group of related proteins that regulate apoptosis in a variety of ways. The success of anticancer treatments often hinges on the ability to induce cancer cell death by apoptosis. As a result, there has been a great deal of interest in developing drugs that can inhibit the antiapoptotic members of the BCL-2 pathway. Unfortunately, some of these drugs are also associated with dose-limiting hematologic toxicities, such as neutropenia.
Now, Leverson et al. have used a toolkit of BCL-2 family inhibitors with different specificities to show that specifically
inhibiting BCL-X L (one member of this protein family) is effective for killing tumors, but without the common side effects seen with less selective drugs.

 

 
A complete electronic version of this article and other services, including high-resolution figures, can be found at:

http://stm.sciencemag.org/content/7/279/279ra40.full.html

Supplementary Material can be found in the online version of this article at: http://stm.sciencemag.org/content/suppl/2015/03/16/7.279.279ra40.DC1.html
Information about obtaining reprints of this article or about obtaining permission to reproduce this
article in whole or in part can be found at:

http://www.sciencemag.org/about/permissions.dtl

 

 

 

 

 

 

 

 

 

 

 

 

 
Science Translational Medicine (print ISSN 1946-6234; online ISSN 1946-6242) is published weekly, except the last week in December, by the American Association for the Advancement of Science, 1200 New York Avenue NW, Washington, DC 20005. Copyright 2015 by the American Association for the Advancement of Science; all
rights reserved. The title Science Translational Medicine is a registered trademark of AAAS.