<<

Published OnlineFirst May 2, 2019; DOI: 10.1158/2159-8290.CD-19-0125

RESEARCH ARTICLE

The TP53 Apoptotic Network Is a Primary Mediator of Resistance to BCL2 Inhibition in AML Cells

Tamilla Nechiporuk1,2, Stephen E. Kurtz1,2, Olga Nikolova2,3, Tingting Liu1,2, Courtney L. Jones4, Angelo D’Alessandro5, Rachel Culp-Hill5, Amanda d’Almeida1,2, Sunil K. Joshi1,2, Mara Rosenberg1,2, Cristina E. Tognon1,2,6, Alexey V. Danilov1,2, Brian J. Druker1,2,6, Bill H. Chang2,7, Shannon K. McWeeney2,8, and Jeffrey W. Tyner1,2,9

ABSTRACT To study mechanisms underlying resistance to the BCL2 inhibitor in acute myeloid (AML), we used a genome-wide CRISPR/Cas9 screen to identify knockouts resulting in drug resistance. We validated TP53, BAX, and PMAIP1 as whose inactivation results in venetoclax resistance in AML cell lines. Resistance to venetoclax resulted from an inability to execute apoptosis driven by BAX loss, decreased expression of BCL2, and/or reliance on alternative BCL2 family members such as BCL2L1. The resistance was accompanied by changes in mitochondrial homeostasis and cellular metabolism. Evaluation of TP53 knockout cells for sensitivities to a panel of small-molecule inhibitors revealed a gain of sensitivity to TRK inhibitors. We relate these observations to patient drug responses and gene expression in the Beat AML dataset. Our results implicate TP53, the apoptotic network, and mitochondrial functionality as drivers of venetoclax response in AML and suggest strategies to overcome resistance.

SIGNIFICANCE: AML is challenging to treat due to its heterogeneity, and single-agent therapies have universally failed, prompting a need for innovative drug combinations. We used a genetic approach to identify genes whose inactivation contributes to drug resistance as a means of forming preferred drug combinations to improve AML treatment.

See related commentary by Savona and Rathmell, p. 831.

1Division of Hematology and Medical Oncology, Oregon Health and Sci- Note: Supplementary data for this article are available at Cancer Discovery ence University, Portland, Oregon. 2Knight Cancer Institute, Oregon Health Online (http://cancerdiscovery.aacrjournals.org/). 3 and Science University, Portland, Oregon. Department of Biomedical Corresponding Author: Jeffrey W. Tyner, Oregon Health and Science Engineering, Oregon Health and Science University, Portland, Oregon. University, 3181 SW Sam Jackson Park Road, Mailcode KR-HEM, Port- 4 Division of Hematology, University of Colorado Denver, Aurora, Colorado. land, Oregon 97239. Phone: 503-346-0603; Fax: 503-494-3688; E-mail: 5 Department of Biochemistry and Molecular Genetics, University of Colorado [email protected]. Denver, Aurora, Colorado. 6Howard Hughes Medical Institute, Oregon Health and Science University, Portland, Oregon. 7Department of Pediatrics, Oregon Cancer Discov 2019;9:1–16 Health and Science University, Portland, Oregon. 8Division of Bioinformatics doi: 10.1158/2159-8290.CD-19-0125 and Computational Biology, Department of Medical Informatics and ©2019 American Association for Cancer Research. Clinical Epidemiology, Oregon Health and Science University, Portland, Oregon. 9Department of Cell, Developmental and Cancer Biology, Knight Cancer Institute, Oregon Health and Science University, Portland, Oregon.

OF1 | CANCER DISCOVERY JULY 2019 www.aacrjournals.org

Downloaded from cancerdiscovery.aacrjournals.org on September 29, 2021. © 2019 American Association for Cancer Research. Published OnlineFirst May 2, 2019; DOI: 10.1158/2159-8290.CD-19-0125

INTRODUCTION (13–17). The latest generation of BH3 mimetics, venetoclax, selectively targets BCL2, thus overcoming the deficiencies Resisting cell death is a hallmark of cancer as well as an and side effects of its predecessors, navitoclax and ABT-737, essential feature of acquired drug resistance (1–3). To main­ which targeted BCL2 and BCL2L1. Selective MCL1 inhibitors tain their growth and survival, cancer cells often modulate cell have also been developed and are in clinical development death mechanisms by overexpression of antiapoptotic BCL2 (18–20). Venetoclax has FDA approval for use in patients family members, including BCL2, BCL2L1, and MCL1. This with chronic lymphocytic leukemia (CLL), and its efficacy overexpression aims to neutralize BH3 domains of the apoptotic has been evaluated for other hematologic malignancies; it activators BID, BIM, and PUMA, thus preventing liberation of has been recently approved in combination with a hypo­ the proapoptotic BAX and BAK from inhibitory inter­ methylating agent for (AML; refs. actions with antiapoptotic BCL2 family proteins (4–6). Elevated 21–25). Intrinsic or acquired resistance is a central concern levels of MCL1 are commonly observed in cancer and are with single-agent targeted therapies, with many instances associated with poor survival in myeloid hematologic malignan­ now documenting favorable initial responses that give way to cies (7–10). Aberrant prosurvival regulation can also be achieved disease relapse resulting from loss of drug sensitivities (26). by inactivation of function, which controls many aspects For venetoclax, both acquired and intrinsic resistance is of apoptosis (11, 12). These processes alter the balance of pro­ anticipated from initial clinical studies using venetoclax apoptotic and antiapoptotic proteins, permitting cancer cells, mono­therapy on patients with relapsed/refractory AML, otherwise primed for apoptosis, to survive and resist therapies. which have shown modest efficacy due to either intrinsic or The elucidation of BH3 protein family interactions acquired drug resistance (24). The efficacy of venetoclax in involved in promoting or inhibiting apoptosis has enabled patients with AML is dependent on BCL2 expression levels, the development of small-molecule inhibitors targeting spe­ and resistance to venetoclax may result from overexpression of cific antiapoptotic family members leading to release of the the antiapoptotic proteins MCL1 or BCL2L1 (21, 24). Similar proapoptotic proteins BAX and BAK, their oligomerization mechanisms have been observed in cell lines, as and mitochondrial outer membrane permeabilization, and several studies have shown resistance to venetoclax developing ultimately execution of the irreversible apoptotic cascade primarily through alterations in levels of other antiapoptotic

JULY 2019 CANCER DISCOVERY | OF2

Downloaded from cancerdiscovery.aacrjournals.org on September 29, 2021. © 2019 American Association for Cancer Research. Published OnlineFirst May 2, 2019; DOI: 10.1158/2159-8290.CD-19-0125

RESEARCH ARTICLE Nechiporuk et al. proteins (27, 28). AML cell lines rendered resistant to veneto­ Fig. 1E). These findings are concordant with results of a simi­ clax through prolonged exposure develop a dependency on lar screen, described in a companion paper by Chen and col­ MCL1 or BCL2L1 (19, 29). In AML cells, the dependency on leagues in this issue. To confirm our results, we performed a MCL1 for survival can be overcome by MAPK/GSK3 signaling parallel screen in MOLM-13 cells using a distinct genome- via activation of p53, which when coupled with BCL2 inhibi­ wide library, Brunello (33), which also identifiedTP53 , BAX, tion enhances the antileukemic efficacy of venetoclax (30). and PMAIP1 among the top-enriched sgRNAs (Fig. 1F; Sup­ Drug combinations offer a strategy to overcome limitations plementary Tables S1 and S2). The distribution of concord­ in the durability and overall effectiveness of targeted agents ance, defined as a statistically significant increase or decrease resulting from intrinsic mechanisms of resistance. Understand­ (P < 0.05) in the same direction among all sgRNAs for a ing these mechanisms is essential for identifying drug targets given gene, showed higher average percent concordance in that will form the basis for new combinatorial therapeutic the Y. Kosuke library screen compared with the Brunello strategies to circumvent drug resistance. To identify essential library screen (0.77 ± 0.07, n = 916 vs. 0.62 ± 0.004, n = 2310, target genes and pathways contributing to venetoclax resist­ unpaired two-tailed t test, P < 0.0001); therefore, we focused ance in AML, we used a genome-wide CRISPR/Cas9 screen on the Y. Kosuke screen results for pathway analyses. Rank­ on an AML patient–derived cell line, MOLM-13. Our findings ing genes with statistically significant enriched sgRNAs P( < identify TP53, BAX, and PMAIP1 as key genes whose inactiva­ 0.05) based on concordance as described above and a ≥3 log tion establishes venetoclax resistance, centering on regulation fold change compared with controls identified 234 genes of the mitochondrial apoptotic network as a mechanism of (Supplementary Table S3). Functional enrichment analyses controlling drug sensitivity. Cell lines with TP53 or BAX knock­ of these 234 candidates indicated an enrichment in genes outs (KO) show perturbations in metabolic profile, energy involved in mitochondrial processes, such as apoptosis tran­ production, and mitochondrial homeostasis. In addition, TP53 scriptional regulation (TP53 and TFDP1; refs. 13 and 34) KO cells acquired sensitivity to TRK inhibitors, suggesting a and apoptosis sensitization [PMAIP1 (NOXA); ref. 35], genes new dependency on this survival pathway, indicating changes required for mitochondrial membrane pore formation (BAX, in transcriptional activity in the venetoclax-resistant setting BAK, SLC25A6, and TMEM14A; refs. 13, 34), and additional may offer drug combinations to overcome resistance. genes involved in apoptosis (proapoptotic BNIP3L; ref. 36; Supplementary Table S4). The top-enriched sgRNAs showed RESULTS a high degree of concordance in fold change relative to con­ trols (Fig. 1G; Supplementary Tables S1 and S2) and had Genome-Wide Screen Reveals p53 Apoptosis a baseline of more than 100 normalized counts in control Network Controlling Venetoclax Resistance samples, precluding artificial inflation of fold changes in To understand possible mechanisms of venetoclax resist­ resistant cells (Fig. 1H). ance in the setting of AML, we screened an AML cell line, MOLM-13, which harbors the most common genetic lesion Prioritization of Genome-Wide Screen Candidates observed in AML (FLT3-ITD), for loss-of-function genes Our study used two independent sgRNA guide libraries, conferring venetoclax resistance. We initially generated Cas9- which provided a high degree of confidence with respect to expressing MOLM-13 cells as described (ref. 31; Fig. 1A) the top hits identified. Analysis of genome-wide CRISPR and confirmed Cas9 functionality (Fig. 1B). Systematic screen knockouts is challenged by off-targeting, sgRNA guide genetic perturbations on a genome-wide scale were intro­ efficiency, and other factors that can lead to library-specific duced by infecting Cas9-expressing cells with lentiviruses artifacts and striking differences between libraries (31, 33). carrying a CRISPR library consisting of an average of 5 single- To prioritize candidates for validation, we developed a tier guide RNAs (sgRNA) per gene for 18,010 genes, hereafter structure that incorporates three key factors: evidence (deter­ referred to as Y. Kosuke (31). After transduction, we further mined by the number of sgRNA guide hits per gene), con­ selected cells with puromycin for 5 days to ensure stable cordance (indicated by the agreement across the set of guides guide integration, and then subjected cells to treatment for a given gene), and discovery (based on expanding effect with either 1 μmol/L venetoclax or vehicle alone (DMSO) size threshold) to rank sgRNA hits and enable a progression for 14 days while collecting DNA at various time points to pathway analysis for lower-scoring hits (Supplementary (Fig. 1C). By day 7 of drug treatment, venetoclax reduced Methods). Using this prioritization scheme, the tier 1 hits cell viability by approximately 50% relative to 75% reduction (n = 149) revealed significant biological identity with theTP53 without drug exposure. PCR-amplified libraries of barcodes Regulation of cytochrome C release pathway (Reactome; representing unique sgRNA sequences obtained from DNA corrected P < 0.001), which is concordant with our initial extracted from either control (DMSO) or drug-treated cells analysis. were subjected to deep sequencing and analyzed using a MAGeCK pipeline analysis (ref. 32; Fig. 1D; Supplementary Inactivation of Genes as Single Knockouts Tables S1 and S2). Plotting the average log fold change for Confirms Resistance to Venetoclax and sgRNA counts per gene versus cumulative P values generated Validates the Screen by robust rank aggregation analyses revealed a high enrich­ To validate the screen hits, we designed several individual ment of p53 and BAX in venetoclax-resistant populations sgRNAs to knock out TP53, BAX, PMAIP1, TFDP1, and several as well as a significant enrichment for several other genes other top candidate genes along with nontargeting controls. involved in the apoptosis pathway or mitochondrial homeo­ Analyses of drug sensitivity at 14 days after transduction of stasis, such as PMAIP1 (NOXA), TFDP1, and BAK1 (BAK; MOLM-13 cells with individual sgRNAs revealed a loss of

OF3 | CANCER DISCOVERY JULY 2019 www.aacrjournals.org

Downloaded from cancerdiscovery.aacrjournals.org on September 29, 2021. © 2019 American Association for Cancer Research. Published OnlineFirst May 2, 2019; DOI: 10.1158/2159-8290.CD-19-0125

TP53 Network Mediates Resistance to BCL2 Inhibition RESEARCH ARTICLE

A B 250k 5 Q1 Q2 10 FSC-A SSC-A subset 0.710% 8.03% EF1a Cas9 2A Blsd 200k 86.8% 104 150k Empty 103 ssc-A 100k 561-2-A sgRNA

u6 Scaff. PGK GFP 2A Cherry 2 50k 10 Empty 0 Q4 Q3 GFP 0 86.6% 2.63% 0 50K 100K 150K 200K 250K 0102 103 104 105 FSC-A 488-1-A

250k Q1 Q2 105 FSC-A SSC-A subset 3.25% 1.47% 200k 88.2% 104 150k GFP 103

100k 561-2-A sgRNA

2 50k 10 Q4 Q3 0 94.0% 1.26% 0 0 50K 100K 150K 200K 250K 0102 103 104 105 Cherry FSC-A 488-1-A C Infect Select Collect Days 0–14 GFP

Puro Ven PCR amplify Deep Analyze DMSO barcodes sequence enrichment/ dropout

sgRNA Cas9+ library MOLM13 D EF

) 7 TP53

2 8 20 6 BAX 7 BAX

10 5 10 6 TP53 10 4 5 TFDP1 4 PMAIP1

P , − log P , − log AXIN2 BNIP3L . counts + 1 (log 5.0 PMAIP1 AXIN2

No rm 0 BAK1 −10 −8 −6 −4 −2 0 246810 −12 −8 −404812 16 Day 0Day 7Day 14Day 21Day 7Day 14 Fold change (treated/untreated), Fold change (treated/untreated), DMSO Ven log2 log2 GH 10 20 2 ) 2 5 15 10 0 5

−5 0 Fold change (log Fold Norm. counts, log counts, Norm.

−5 −10 FAU C4A AIP BAXTP53 FADD BAX FAU C4A AIP CCL18 TFDP1 TP53 FADD MAD2L2 PMAIP1 PDCD10EPHA7 CCL18 TFDP1 EPHA7 PYCARD MAD2L2 PMAIP1 PYCARDPDCD10

Figure 1. Genome-wide CRISPR/Cas9 screen in AML cells identifiedTP53, BAX, and other apoptosis network genes conferring sensitivity to vene- toclax. A, Schematic representation of lentiviral vectors described elsewhere in detail (31) and used for the delivery and functional assay of Cas9. Top, Cas9-expressing vector; Blsd, blasticidin selection gene; EF1a, intron-containing human elongation factor 1a promoter; Cas9, codon-optimized Strepto- coccus pyogenes double-NLS-tagged Cas9; 2A, Thosea asigna virus 2A peptides. Bottom, vector carrying dual fluorescent proteins; GFP and mCherry expressed from the PGK promoter; U6 denotes human U6 promoter driving GFP sgRNAs or empty cassette; Scaff denotes sgRNA scaffold. B, Functional assay for Cas9 activity in MOLM-13 cells transduced with virus carrying an empty sgRNA cassette (top) or sgRNA targeting GFP (bottom), assessed by flow cytometry 5 days after transduction. Note the significant decrease in GFP signal in the presence of sgRNA targeting GFP. C, Schematic representation of genome-wide screen for drug resistance. The sgRNA library (31) was transduced into Cas9-expressing MOLM-13 cells, selected with puromycin for the integration of sgRNA-carrying virus for 5 days and DNA collected from cells exposed to venetoclax (1 μmol/L) or vehicle (DMSO) for various time points (days 0, 7, 14, and 21). sgRNA barcodes were PCR-amplified and subjected to deep sequencing to analyze for enrichment and/or dropout. D, Normalized counts of sgRNAs from collected DNA samples, median, upper, and lower quartiles are shown for representative replicate samples. E, F, Enrichment effect in Y. Kosuke (E) and Brunello (F) library screens for loss of sensitivity to venetoclax. Fold change and corresponding P values are plotted; genes represent- ing significant hits in both libraries are highlighted in red. G, Enrichment extent plotted as fold change over control following venetoclax exposure (day 14) for the set of individual top hit sgRNAs per gene is shown (Y. Kosuke library). H, Box and whisker plots spanning min/max values of normalized counts for control (left boxes in each pair) and venetoclax treatment (right boxes in each pair) combined for all sgRNAs per gene. Top hits are shown.

JULY 2019 CANCER DISCOVERY | OF4

Downloaded from cancerdiscovery.aacrjournals.org on September 29, 2021. © 2019 American Association for Cancer Research. Published OnlineFirst May 2, 2019; DOI: 10.1158/2159-8290.CD-19-0125

RESEARCH ARTICLE Nechiporuk et al. venetoclax sensitivity (Fig. 2A). The top candidates, including Inactivation of TP53 and BAX Diminishes the TP53 and BAX, were also validated by single-guide inactivation Apoptotic Response to Venetoclax in an additional cell line, MV4-11 (Fig. 2B and C), with many To characterize venetoclax sensitivity in MOLM-13 cells IC50 values significantly exceeding initial drug concentrations with KOs for TP53, BAX, or PMAIP1, we evaluated an early used for the sgRNA screen. Analyses of protein levels for the marker of apoptosis by measuring the percentage of Annexin top candidates, BAX, TP53, and PMAIP1, demonstrated sig­ V–positive cells in response to venetoclax treatment (100 nmol/L) nificant loss of protein upon sgRNA inactivation (Fig. 2D; (41). Cells with KOs for TP53, BAX, or PMAIP1 showed a sig­ Supplementary Fig. S1A and S1B). Although BAX is reported nificant decrease in the number of cells undergoing apoptosis to be a p53 transcriptional target (reviewed in ref. 37), its levels (Fig. 4A and B). This was accompanied by sustained phospho­ remained unchanged when TP53 was inactivated, indicating rylation of MAPK1/3 (ERK1/2) during venetoclax treatment that other transcriptional factors may regulate BAX levels in and increase in total levels of AKT and MAPK1/3 in cells with these cells (38). Levels of other p53 target gene products such inactivation of TP53 and BAX (Fig. 4C). as PMAIP1, PUMA, and BAK1 were decreased in TP53 KO cells (Supplementary Fig. S1A and S1C). At the same time, levels of antiapoptotic proteins BCL2 and MCL1 were reduced in Resistance to Venetoclax Coincides with all four tested TP53 knockout lines, inversely correlating with Protection from Mitochondrial Stress and increased BCLXL expression (Fig. 2D; Supplementary Fig. Increases in Oxidative Phosphorylation S1C). Analysis of protein levels revealed increases in the ratios The top hits from our CRISPR screen pointed toward per­ of BCLXL to BCL2 in TP53 KO cells (Supplementary Fig. turbations in mitochondrial regulators and effectors of the S1D). Venetoclax directly binds BCL2; therefore, a decrease apoptotic program. p53, BAX, PMAIP1, TFDP1, and BAK in BCL2 expression in TP53 KO cells may contribute to the function in a proapoptotic manner: p53 acts a sensor of loss of sensitivity. The inverse correlation of low BCL2/high cellular stress and transcriptional regulator of several proa­ BCLXL expression with null TP53 status is similarly observed poptotic genes; TFDP1 activates and translocates PMAIP1 in pediatric acute lymphoblastic leukemia (39). In a large AML into the mitochondria; PMAIP1 promotes activation of the patient cohort (Beat AML; ref. 40), BCL2 expression positively apoptotic program; and BAX/BAK function as its effectors. correlated with TP53 expression, whereas both BCLXL and In this manner, we observed that inactivation of TP53 led MCL1 had inverse correlations (Fig. 2E). Increased levels of to alterations in the expression of antiapoptotic BCL2 fam­ BCLXL in TP53 KO cells correlate with sensitivity to a BCL2/ ily members and many TP53 transcriptional targets that BCLXL inhibitor (AZD-4320; Fig. 2F). Because both MOLM- affect mitochondrial homeostasis and cellular metabolism 13 and MV4-11 cell lines carry an FLT-ITD mutation, we (Fig. 2D; Supplementary Fig. S1). TP53 inactivation also introduced loss-of-function mutations in TP53, BAX, and altered response to a variety of stress stimuli and rendered PMAIP1 into OCI-AML2, an FLT3 wild-type (WT) cell line, cells resistant to many other pharmacologic agents. These and observed similarly diminished sensitivities to venetoclax changes prompted us to investigate the impact of overall (Supplementary Fig. S1E and S1F). response to mitochondrial stress in cells briefly exposed to venetoclax (Fig. 5A–D). Cells with inactivation of p53 and AML Patient Samples with Mutant TP53 Status BAX were less likely to undergo mitophagy in response to the or Low Expression Levels of TP53 and BAX Have mitochondrial uncoupler carboyl cyanide m-chlorophenyl Diminished Sensitivity to Venetoclax Ex Vivo hydrazine (CCCP), regardless of venetoclax treatment, with To assess whether the top candidates from the screen more pronounced protection observed in TP53 mutants, have a direct relevance to gene perturbation in patients suggesting overall changes in mitochondrial homeostasis with AML, we analyzed the correlation of venetoclax drug (Fig. 5A and B). Additionally, TP53- and BAX-mutant cells sensitivity with gene-expression levels and mutational sta­ were less likely to lose mitochondrial membrane potential tus using the Beat AML patient sample dataset (40). Our in response to CCCP alone, with venetoclax potentiating the analyses indicated that 16 patient samples harboring del­ CCCP uncoupling action (Fig. 5C and D). eterious TP53 mutations showed a trend toward reduced Venetoclax inhibits BCL2 by mimicking the BH3 bind­ venetoclax sensitivity (measured as increases in AUC in a ing of apoptotic sensitizers/activators. Both venetoclax and nonlinear modeling of drug response curves) compared with chemical inhibition of BCL2 influence energy production by 282 AML patient samples with WT TP53 (Mann–Whitney decreasing oxidative phosphorylation in the mitochondria t test, P < 0.04; Fig. 3A). Furthermore, separation of TP53 of leukemia stem cells (42) and can affect various metabolic expression levels into high and low quartiles across the activities in the cells, including amino acid catabolism (43). total TP53 WT cohort showed high quartile samples were In addition, we observed reduced levels of PUMA in TP53 more sensitive to venetoclax than low quartile samples KO cells (Supplementary Fig. S1C), recently identified as a (Fig. 3B). Regardless of TP53 mutational status, both TP53 p53-induced metabolic switch (44). Because p53 regulates and BAX expression levels were inversely correlated with vene­ many genes whose products determine the balance of reac­ toclax sensitivity (high AUC), underscoring the relevance of tive oxygen species (ROS) production in response to cellular our CRISPR-derived loss-of-function candidates to patient- stress (45–47), we evaluated whether oxidative phosphoryla­ level responses ex vivo (Fig. 3C). As expected, BCL2 levels tion was perturbed in cells inactivated for TP53 by measuring showed an inverse correlation with venetoclax sensitivity, rates of oxygen consumption and electron transport across whereas BCLXL expression directly correlated with veneto­ the mitochondrial membrane. Both outputs were enhanced clax sensitivity (Fig. 3C). in mutant cells (Fig. 5E and F), accompanied by a moderate

OF5 | CANCER DISCOVERY JULY 2019 www.aacrjournals.org

Downloaded from cancerdiscovery.aacrjournals.org on September 29, 2021. © 2019 American Association for Cancer Research. Published OnlineFirst May 2, 2019; DOI: 10.1158/2159-8290.CD-19-0125

TP53 Network Mediates Resistance to BCL2 Inhibition RESEARCH ARTICLE

A 150 B 150 MV4-11P MOLM-13P 100 NT_g 100 NT_g TP53_g TP53_g

% Viable % 50 BAX_g Viable % 50 BAX_g PMAIP1_g FAU_g

010 100 1,000 10,000 0 10 1001,000 10,000 Venetoclax (nmol/L) Venetoclax (nmol/L)

D C MOLM-13 MV4-11 104 MOLM-13PNT_g BAX_g KMT2C_gTP53_g1TP53_g2TP53_g3TP53_g4 TP53 103

102 BAX (nmol/L) 50 101 BCL2 IC

1 BCLXL

NT_g MCL1 FAU_g TP53_gBAX_g Parental BCL11_g TFDP1_g PMAIP1_g EPHA7_g PDC10_g MAD2L2_g PYCARD_g GAPDH

E k = 0.33, r = 0.35 k = 0.45, r = 0.15 k = −0.29, r = −0.05 k = −0.73, r = −0.05 10 10 **** * 10 **** 10 **** 8 8 8 8 6 6 6 6 4 4 4 4 2 2 2

TP53, RPKM 2 TP53, RPKM TP53, RPKM TP53, RPKM 0 0 0 0 5678910 5678910 0510 15 6810 12 14 BCL2, RPKM BAX, RPKM BCLXL, RPKM MCL1, RPKM

F 150 150 NT_g NT_g e e 100 100 TP53_g TP53_g PMAIP1_g PMAIP1_g

50 Via bl % 50 % Via bl % BAX_g BAX_g 0 0 010 100 1,000 10,000 010100 1,00010,000 Venetoclax (nmol/L) AZD-4320 (nmol/L)

Figure 2. A, Confirmation of genes conferring venetoclax resistance in MOLM-13 and MV4-11 cells and correlations with changes in apoptotic gene transcription in patients with AML. MOLM-13 (A) or MV4-11 (B) cells were transduced with lentiviruses carrying single sgRNA/Cas9 constructs targeting TP53, BAX, PMAIP1, FAU, or control (nontargeting; NT_g). Ten days after transduction, venetoclax sensitivity was measured in triplicate by colorimetric MTS assay using a 7-point concentration range (10 nmol/L to 10 μmol/L). Percentages of viable cells, after normalization to untreated controls, were fit using nonlinear regression analyses; mean and standard errors are shown. MOLM-13P, parental MOLM-13 cells; MV4-11P, parental MV4-11 cells.

C, Histogram (log scale) summarizing IC50 estimates from triplicate measurements for venetoclax sensitivity in parental MOLM-13 (black shaded), parental MV4-11 cells (gray shaded), or cells transduced with indicated sgRNA/Cas9 viruses. D, Western blot analyses of proteins extracted from MOLM-13 cells, transduced with indicated sgRNA/Cas9 viruses, and identified with antisera to BAX, BCL2, BCLXL, MCL1, and GAPDH. Note that forTP53 , four sgRNAs producing distinct knockout alleles were used to confirm changes in expression levels of BCL2, BCLXL, and MCL1. sgRNA targetingKMT2C was used as an additional unrelated sgRNA control. E, Correlation of expression of TP53 and selected genes in an AML patient sample cohort (n = 246; ref. 40). k, slopes generated by linear regression; r, Spearman coefficient; *,P < 0.05; ****, P < 0.0001. F. Sensitivity profiles of venetoclax (left) and AZD-4320 (right) on MOLM-13 cells transduced with lentiviruses carrying single sgRNA/Cas9 constructs targeting TP53, BAX, PMAIP1, and NT (nontargeting) control, measured as in A using a 7-point concentration range (8 nmol/L to 10 μmol/L).

increase in cellular ROS (Fig. 5G), as oxidative phosphoryla­ that apoptosis can still be triggered in venetoclax-resistant tion is a major cellular source of ROS production (48). Inter­ cells through disruption of mitochondrial homeostasis estingly, both TP53 and BAX KO cells were still susceptible to (Fig. 5H). These results suggest an inhibition of a general cell death by elesclomol, a drug that induces cellular death mitochondrial stress response coupled with increased cel­ through an increase in cellular oxidation (49), suggesting lular respiration.

JULY 2019 CANCER DISCOVERY | OF6

Downloaded from cancerdiscovery.aacrjournals.org on September 29, 2021. © 2019 American Association for Cancer Research. Published OnlineFirst May 2, 2019; DOI: 10.1158/2159-8290.CD-19-0125

RESEARCH ARTICLE Nechiporuk et al.

A B 400 *** 400 LOF C C GOF *** *** * 300 Splice 300 WT

200 200

100 100 VEN sensitivity, AU VEN sensitivity, VEN sensitivity, AU VEN sensitivity,

0 0

mut WT LE HE total WT WT WT TP53 TP53 TP53 TP53 TP53 C k = −20.13, r = −0.31 k = −26.09, r = −0.14 k = −34.79, r = −0.20 k = 4.3, ns 400 400 400 400 **** **** ****

C 300 C 300 C 300 C 300 200 200 200 200

VEN AU 100 VEN AU 100 VEN AU 100 VEN AU 100 0 0 0 0 0246810 567 89 0246 81002 46810 TP53, RPKM BAX, RPKM BCL2, RPKM PMAIP1, RPKM

k = 20.26, r = 0.31 k = 26.87, r = 0.27 k = 24.5, r = 0.63 k = 0, ns 400 400 400 400 **** **** ****

300 C

C 300 C 300 300 C 200 200 200 200 VEN AU

VEN AU VEN AU 100 100 100 VEN AU 100 0 0 0 0 02468 6 8 10 12 14 0510 15 45678 BAK1, RPKM BBC3, RPKM MCL1, RPKM BCLXL, RPKM

Figure 3. Venetoclax (VEN) sensitivity in patients with AML inversely correlates with TP53 mutations and low expression of TP53 and BAX. A, Com- parison of VEN sensitivity between TP53-mutant (TP53mut; n = 16) and TP53 WT (TP53WT; n = 282) patient samples (Mann–Whitney, two-tailed). Within TP53-mutant groups, circles, triangles and square symbols denote loss of function (LOF), gain of function (GOF), and splice variants correspondingly. B, Comparison of VEN sensitivity between the lowest expressing (LE; n = 71) and highest expressing (HE; n = 71) quartiles for TP53 in WT patient sam- ples. Median expression levels for both LE and HE groups differ significantly from the median expression levels in all samples (ANOVA with Tukey post-test). C, Correlation between gene-expression levels from AML patient samples for the indicated genes and venetoclax sensitivities represented by AUC (n = 246; ref. 40). k, slope values generated by linear regression; r, Spearman coefficient. *,P < 0.05; ***, P < 0.001 ****; P < 0.0001; ns, not significant.

Perturbations in the Metabolic Profile of tary Fig. S2; Supplementary Table S5). Additionally, a decrease Venetoclax-Resistant Cells in amino acid levels and urea cycle intermediates highlighted the most significantly affected pathways (Fig. 6A–D) in both Unlike the Warburg effect where rapidly proliferating cancer BAX- and TP53-mutant cells in comparison with control cells. cells in glucose-rich conditions switch to less efficient aerobic These reductions were accompanied by increases in nucleo­ glycolysis using the pentose phosphate shunt (48), increased tide synthesis, suggesting the utilization of more glucose and oxygen consumption in TP53 KO cells suggests increased amino acids for the purpose of nucleotide synthesis and cellu­ energy production via anaerobic glycolysis in venetoclax-resist­ lar proliferation (Fig. 6C and D; Supplementary Table S5). We ant cells. Accordingly, we performed a global metabolomics also performed parallel metabolomic studies on TP53-mutant analysis of MOLM-13 cells with inactivation of TP53, BAX, and and WT patient samples (Supplementary Fig. S3A–S3C; Sup­ nontargeting controls. Unexpectedly, we observed a significant plementary Table S5) and found concordance for decreases in decrease in major glycolysis intermediates, such as D-glucose glucose and pyruvate levels, amino acids, and urea cycle inter­ and D-glucose-6-phosphate; however, levels of pyruvate, the mediates (Supplementary Fig. S3A, 3B, 3D, and 3E), offering final metabolite entering into the tricarboxylic acid (TCA) validation of the altered metabolic states observed in the cell cycle, remained unperturbed (Fig. 6A and B; Supplementary line models (Supplementary Fig. S3). These results, together Fig. S2; Supplementary Table S5). We also observed a decrease with the diminished sensitivity to venetoclax and perturba­ in levels of 6-phospho-D-gluconate, but not in subsequent tions in mitochondrial responses, illustrate the significance of intermediates in the pentose phosphate pathway (Supplemen­ our screen hits as depicted in Fig. 6E. From the transcriptional

OF7 | CANCER DISCOVERY JULY 2019 www.aacrjournals.org

Downloaded from cancerdiscovery.aacrjournals.org on September 29, 2021. © 2019 American Association for Cancer Research. Published OnlineFirst May 2, 2019; DOI: 10.1158/2159-8290.CD-19-0125

TP53 Network Mediates Resistance to BCL2 Inhibition RESEARCH ARTICLE

A Alteration of Drug Sensitivities in TP53 100 24 h 48 h and BAX KO Cells 80 72 h cells + 60 To evaluate whether the mitochondrial and metabolic changes led to alterations in additional signaling pathways, 40 we evaluated drug sensitivities in MOLM-13 and MV4-11 cells 20 with TP53 and BAX KOs relative to parental and nontargeting

% Annexin V % Annexin 0 control cells using a panel of small-molecule inhibitors that MOLM-13P NT_g TP53_g BAX_g target a variety of distinct signaling pathways activated in VEN −+−+−+ −+ cancers. We observed many commonly altered drug responses 100 nmol/L among venetoclax-resistant cell lines, including a core of B drug responses that were common to both modified MOLM- 100 DMSO 13 and MV4-11 cells (Fig. 7A; Supplementary Table S6). In 80 VEN 48 h 50 nmol/L addition to loss of venetoclax sensitivities, tyrosine kinase cells + VEN 48 h 100 nmol/L inhibitor sensitivities to the AKT/PI3K pathway (GDC-0941, 60 PI-103), as well as YM-155, an inhibitor of survivin, were 40 decreased (Fig. 7A; Supplementary Table S6). Although the 20 BAX KO showed a similar sensitivity to the MCL1 inhibi­

% Annexin V % Annexin 0 tor AZD5991 relative to control cells, AZD5991 sensitivity _g was diminished in TP53 KO MOLM-13 cells, indicating that NT_g these cells may not depend on MCL1 for survival. The MCL1 TP53_g dependency further decreased in prolonged TP53 KO cul­ MOLM-13P PMAIP1 tures, possibly due to compensatory mechanisms of survival. C DMSO VEN 100 nmol/L Except for venetoclax, few, if any, inhibitors showed concord­ pERK1/2 ant changes between BAX-inactivated MV4-11 and MOLM-13 cells (Supplementary Table S6). Additional changes for cells ERK1/2 with TP53 inactivation included the loss of sensitivity to all six tested FLT3 inhibitors (crenolanib, , , pAKT , , and gilteritinib; Fig. 7B; Supple­ mentary Fig. S4A). Cells with TP53 KO showed sustained AKT gain of sensitivity to several NTRK/ALK/ROS1 inhibitors, namely, (50), (51, 52), GW-2580 (53), GAPDH and AZD1480 (ref. 54; Fig. 7A, C and D; Supplementary Fig. S4B; Supplementary Table S6). Neurotropin receptors (NTRK1/2/3, also known as TRKA/B/C) are well documented

NT_g NT_g to play a role in neuronal development. NTRK receptors bind BAX_g BAX_g TP53_g TP53_g neurotrophin ligands and are single transmembrane catalytic MOLM-13P MOLM-13P receptors with intracellular tyrosine kinase activity. NTRK receptors have recently gained attention for their oncogenic role in a variety of cancer types, occurring as a component Figure 4. Cells with loss-of-function alleles for TP53, PMAIP1, or of fusion genes (55–58). NTRK receptor activation is cou­ BAX have diminished apoptosis in response to venetoclax treatment. A and B, Sensitivities to venetoclax in MOLM-13 parental cells (MOLM- pled to RAS/MAPK, PI3K, or PLCγ signaling pathways to 13P) and MOLM-13 cells with sgRNA-inactivated alleles, as indicated, were promote survival and proliferation (58, 59). Although some assessed by flow cytometry after staining with the early apoptosis marker, TRK inhibitors have additional targets beyond NTRKs, such Annexin V, following 24, 48, and 72 hours of exposure to venetoclax. Histogram represents mean and standard deviation for three replicates of as ALK and ROS1, additional testing with a more selective percentage of Annexin V+ cells in the total cell population. C, Western blot TRK inhibitor, , also revealed a profound gain of analysis of proteins extracted from MOLM-13 parental and MOLM-13 cells sensitivity, indicating new dependences on the NTRK path­ transduced with indicated sgRNA/Cas9 viruses and treated overnight with 100 nmol/L venetoclax or vehicle (DMSO), and identified with antisera to way for cell survival (Fig. 7C and D). To test whether NTRK phosphorylated ERK1/2 (pERK1/2; Thr202/Tyr204), ERK1/2, phosphoryl- transcripts were upregulated in TP53 knockout cells, we eval­ ated AKT (pAKT; Thr308), AKT, and GAPDH. uated mRNA expression of NTRK1, NTRK2, and NTRK3 in MOLM-13 and MV4-11 cells with and without TP53 deletion. machinery initiated by p53 in response to cellular stress to Although NTRK1 and NTRK2 were expressed at lower levels in various sensitizers such as PMAIP1 and TFDP1 to downstream KO cells, NTRK3 RNA was highly expressed in TP53 KO cells, mitochondrial pore effectors such as BAX, BAK, SLC25A6, and indicating a transcriptional component of NTRK3 upregula­ TMEM14A, our hits demonstrate the central involvement of tion (Fig. 7E). NTRK1 was expressed at a lower level, albeit mitochondria in conferring venetoclax sensitivity. Resistance higher, in control cells, explaining sensitivity to entrectinib, may result from alterations in any of these components, for whereas in TP53 KO cells the mRNA expression switches to example, TP53 mutation affecting BCL2 and BCL2L1 expres­ NTRK3. Analyses of TRK protein levels revealed increases sion (Fig. 6F) or BAX loss of function affecting mitochondrial in overall and phosphorylated levels of TRK proteins in the outer membrane pore formation (Fig. 6G). p53 loss-of-function setting, and we observed significant

JULY 2019 CANCER DISCOVERY | OF8

Downloaded from cancerdiscovery.aacrjournals.org on September 29, 2021. © 2019 American Association for Cancer Research. Published OnlineFirst May 2, 2019; DOI: 10.1158/2159-8290.CD-19-0125

RESEARCH ARTICLE Nechiporuk et al.

A DMSO DMSO + CCCP VEN VEN + CCCP B 44.1 65.4 48.4 87.4 * MOLM-13P 100 *** NT_g ns NT_g 80 *** TP53_g1 TP53_g2 60 ns * BAX_g 40 ns

Mitophagy subset Mitophagy subset Mitophagy subset Mitophagy subset ns 6.6 6.66 10.5 10.5 5.88 5.88 22.2 22.2 20 Mitophagy (%) TP53_g 0

Comp-551-3-A VEN −−++ CCCP −+−+ Crmp-561-3-A Mitophagy dye Lysosome dye D C DMSO DMSO + CCCP VEN VEN + CCCP 120 ns ** MOLM-13P TgTP53_g NT_g 20.8 71.6 43.7 83.5 100 * *** NT_g 80 TP53_g1 60 ns BAX_g2 40 ns 8.96 25.5 12.2 57.3 20 JC-1 monomer (%) 0 VEN −−++ JC1-polymer

JC-1 monomer CCCP −+−+

E F MOLM-13P MOLM-13P NT_g 1,000 NT_g 100 TP53_g1 TP53_g1 Oligomycin FCCP 2-DG Rotenone TP53_g2 Oligomycin FCCP 2-DG Rotenone TP53_g2 500 50 ECAR, mpH/min OCR, pmoles/min 0 0 0 24 52 770 101 0245277 101 Time (min) Time (min) GH 1.5e5 150 MOLM-13P NT_g 0 h BAX_g 1.0e5 100 2 h TP53_g

−5

% Viable % 50 intensity 0.5e

Relative fluorescent Relative 0 0 010 100 1,000 10,000 Elesclomol (nmol/L) NT_g BAX_g TP53_g1TP53_g2TP53_g3 MOLM-13P

Figure 5. Cells with TP53 and BAX inactivation are resistant to mitochondrial stress induced by venetoclax and mitochondrial uncouplers. A, Flow cytometry analyses of MOLM-13 nontargeting and MOLM-13 TP53 KO cells for mitophagy with and without initial 2-hour stimulation with 100 nmol/L venetoclax and/or CCCP. Ten thousand cells were analyzed. Box drawn around doubly stained brighter cells with mitophagy dye is a result of acidifica- tion of the mitophagy dye in lysosomes after fusion with damaged mitochondria. B, Histogram of mitophagy experiments (n = 3). Statistical significance determined by ANOVA with Tukey post-test. Note that the number of mitochondria fused to lysosomes is significantly less in MOLM-13 cells withTP53 knockout (TP53_g) and BAX (BAX_g) relative to control parental and nontargeting (NT_g) cells in the presence of uncoupler CCCP with and without vene- toclax (VEN). *, P < 0.05; **, P = 0.01; ***, P < 0.001; ns, not significant.C . Flow cytometry analyses of mitochondrial depolarization. Percentage of cells with depolarized mitochondria in response to mitochondrial uncoupler CCCP is shown (boxed region: green fluorescent cells representing monomer form of cationic carbocyanine dye JC-1 in depolarized cells). Red fluorescent cells represent polymeric form of JC-1 in hyperpolarized cells. Note protection against depolarization in TP53 and BAX KO cells. D, Histogram of experiments shown in C (n = 3) with 10,000 cells analyzed per sample. Statistical sig- nificance determined by ANOVA with Tukey post-test. *,P < 0.05; **, P = 0.01; ***, P < 0.001; ns, not significant. E, F, Analyses of oxygen consumption rate (OCR) and extracellular acidification rates (ECAR) using Seahorse assay n( = 3). G, Measurement of ROS using 2’,7’-dichlorofluorescin diacetate (DCFDA, Abcam) substrate as an indicator of oxidation inside the cell detected by conversion to fluorescent DCF compound. Note higher rate of ROS inTP53 and BAX KO cells (n = 6). H, Cell viability assay in response to elesclomol. decreases in phosphorylation of TRK and downstream MAPK plementary Fig. S5C). Analysis of cell lysates obtained from signaling in response to treatment with entrectinib in TP53 AML patient samples revealed a marked increase in TRK KO MOLM-13 cells, but not in control or parental cells (Fig. protein with detectable levels of phosphorylated pan-TRK 7F). Treatment of TP53 KO cells with entrectinib resulted in and TRKC in TP53-mutant versus undetectable levels in WT

G1 cell-cycle blockade (Supplementary Fig. S5A and S5B), cases (Fig. 7G). Ex vivo analysis of leukemic blasts from these and treatment of these cells with entrectinib and venetoclax patients with AML showed increased sensitivity to entrec­ together was effective in both parental and TP53 KO cell tinib in those samples harboring TP53 mutations (Supple­ lines but not synergistic relative to each single agent (Sup­ mentary Fig. S6A). A preliminary patient-derived xenograft

OF9 | CANCER DISCOVERY JULY 2019 www.aacrjournals.org

Downloaded from cancerdiscovery.aacrjournals.org on September 29, 2021. © 2019 American Association for Cancer Research. Published OnlineFirst May 2, 2019; DOI: 10.1158/2159-8290.CD-19-0125

TP53 Network Mediates Resistance to BCL2 Inhibition RESEARCH ARTICLE

A BCAlanine, aspartate,

Log10 Log10 Pyrimidine glutamate Class Class 2 15 5′-Phosphoribosyl- 1.0 Serotonin Purine CMP-N-acetylneuram g-Oxalo-crotonate sn-glycero-3-Phosp 0.5 L-serine Amino-acyl t-RNA Xanthine Uracil 1 Tr yptophan L-aspartate 0 L-asparagine PPP 2-Oxoglutaramate Argininosuccinate TCA Arginine and proline 0 10 Inosine −0.5 Adrenic acid 7, 10, Docosapentaenoic a Octadecenoic acid Glycine, serine, and threonine Octadecenoic acid −1 Docosahexaenoic ac Docosahexaenoic ac Tetradecenoic acid −1 Pyruvate Arachidonic acid E Tetradecanoic acid Uracil Hexadecenoic acid − Log(p) 5 Cytosine IMP −2 Diphosphate Arachidonic acid E Malate Glycine Glycine Cystathionine Adenine Xanthine UMP sn-glycero-3-Phosp GMP L-aspartate 0 Fumarate Trans-4-Hydroxy-L- Adenylosuccinic ac Octadecanoic acid 0 0.2 0.4 0.6 0.8 S-adenosyl-L-homoc CMP-N-acetylneuram L-tyrosine Malate Pathway impact (TP53 KO) Cis-p-Coumarate Glutathione disulf N-succinyl-L-citru L-histidine L-leucine/isoleuci L-tryptophan Amino/nucleotide sugar L-phenylalanine L-arginine D PPP Indole-3-acetaldeh Ornithine 8 Starch and sucrose Lactate Cytosine L-histidine Butanoyl-l-carniti L-arginine Creatine Arginine and proline 5-L-glutamyl-L-glu N6-Methyl-L-lysine 6 L-tryptophan Acyl-C4-OH Hydroxy Pyrimidine Indole-3-acetate Acetyl-carnitine a Glycine, threonine, serine 6-phospho-D-glucon Choline Purine N-acetylornithine 5-Hydroxyisourate Creatine Acetylcholine 4 Dimethylglycine L-noradrenaline D-arabitol L-methionine Ornithine − Log(p) Nicotinamide 6-pyruvoyltetrahyd Guanine Phosphocreatine 2 D-glucose 6-phosph L-glutamine 6-phospho-D-glucon D-glucose Dehydroascorbate 5-6-dihydrothymine Phosphocreatine Glycerol 3-phospha D-glucose 0 Ethanolamine phosp UDP Cystathionine 0 0.2 0.4 0.6 0.8 ADP-D-ribose Sedoheptulose 1-ph D-arabitol 2′,3′-cyclic CMP Pathway impact (BAX KO) Ethanolamine phosp F p53 p53 NT_g NT_g BAX_g TP53_g1TP53_g2

E Venetoclax ProsurvivalProapoptotic BCL2 p53 p53*

Transcription BAX/BAK

BH3 TFDP1* NOXA Venetoclax G p53 p53 * BH3 p53 BH3 NOXA MCL1 BCL2 BCL2 BAX/BAK* * BCL2

Venetoclax BCL2 BCL2 ROS BAX/BAK Depolarization

Ub ETC * * ANT BAX/BAK

Mitophagy/autophagy BAX/BAK MOMP TMEM14A Cytochrome C MPTP BAX/BAK TP53 apoptotic network Apoptosis

Figure 6. Metabolic changes in TP53 and BAX KO cells are indicative of increased cell proliferation. A and B, Global metabolomics profile of MOLM-13 TP53 (A) and BAX KO (B) cells. Metabolomic analysis was performed on samples in quadruplicate. Top 50 changed metabolites are shown. C and D, Pathway analysis of metabolites with differential abundance dot plot for MOLM-13 TP53 (C) and BAX KO (D) cells. Red to yellow color gradient indicates higher to lower statistical significance; circle size is proportional to the percentage of affected metabolites within the pathway. E, Summary of genes identified by the CRISPR/Cas9 screen affecting mitochondrial homeostasis, energy production, apoptosis, and venetoclax response. Star symbol indicates identified genes: TP53 (p53), TFDP1 (DP-1), NOXA (PMAIP1), BAX/BAK, TMEM14A, SLC25A6 (ANT3). Mechanisms of venetoclax resistance in cells with inactivation of TP53 (F) or BAX (G). Inactivation of TP53 leads to perturbation in the expression of prosurvival proteins, including BCL2, the primary venetoclax target. Inactivation of BAX leads to inability to build effective mitochondrial outer membrane permeabilization (MOMP) during apoptosis, induced by venetoclax.

JULY 2019 CANCER DISCOVERY | OF10

Downloaded from cancerdiscovery.aacrjournals.org on September 29, 2021. © 2019 American Association for Cancer Research. Published OnlineFirst May 2, 2019; DOI: 10.1158/2159-8290.CD-19-0125

RESEARCH ARTICLE Nechiporuk et al.

AE MOLM-13_TP53_g1 3 *** MOLM-13: ) 10 4 10 MOLM-13_TP53_g2 102 *** Parental 1 MV4-11_TP53_g 10 *** NT_g 2 0 ** 10 TP53_g 10−1 0 MV4-11: 10−2 arb units Parental 10−3 −2 −4 NT_g fold change (log fold 10 NTRK1/2/3 mRNA, 50 10−5 TP53_g

IC −4 NTRK1 NTRK2 NTRK3

PI-103 AT7519 Linifanib YM-155 17-AAG SNS-032 SelinexorNeratinib CrizotinibGW-2580 AZD1480 AZD-5991Volasertib GDC-0941 Venetoclax PalbociclibGDC-0994Barasertib Entrectinib Midostaurin AZ-CDK9I-2 GSK690693 B NVP-ADW742 F 2 10 MOLM-13P 101 NT_g 100 MOLM-13PNT_g TP53_g1TP53_g2MOLM-13PNT_g TP53_g1TP53_g2MOLM-13PNT_g TP53_g1TP53_g2 −1 TP53_g ( µ mol/L) 10 50 10−2 MV4-11P panTRK IC −3 10 MV4-11 pan-pTRK TP53_g pERK1/2

Crenolnib Sorafenib Gilteritinib Venetoclax Quizartinib ERK1/2 Midostaurin C pAKT 102 MOLM-13P AKT 101 100 NT_g GAPDH

( µ mol/L) 1 BAX_g 10− DMSO Ven Entr 50 10−2 TP53_g 100 nmol/L 100 nmol/L IC 10−3

Foretinib GW-2580CrizotinibAZD1480 VenetoclaxEntrectinib Elesclomol Larotrectinib

D 150 MOLM-13P G 17-00325 17-00248 15-00309 17-00779 17-00264 17-00300 pTRKC NT_g1 100 NT_g2 pan-pTRK 50 BAX_g1 % Viable % TP53_g1 panTRK 0 TP53_g2 10 102 103 104 Entrectinib (nmol/L) pERK1/2 150 MOLM-13P ERK1/2 100 NT_g1 NT_g2 GAPDH 50 BAX_g1 % Viable % WT mut TP53_g1 TP53 TP53 0 10 102 103 104 TP53_g2 Larotrectinib (nmol/L)

Figure 7. Cells with loss-of-function alleles for TP53 or BAX have altered sensitivities to small-molecule inhibitors of various signaling pathways. A, MOLM-13 and MV4-11 cells, transduced with indicated sgRNAs/Cas9 viruses, were screened with a panel of inhibitors targeting various molecular path- ways and assessed using the drug-screening method described in Fig. 2A. Fold changes in IC50 values (log10 scale) for those inhibitors concordant across both cell lines relative to nontargeting controls are shown. B, Analyses of FLT3 inhibitor sensitivities, in MOLM-13 and MV4-11 parental and KO cells, as indicated. Drug sensitivity was measured as described in Fig. 2A, with IC50 estimated from triplicates. C, Histogram of sensitivities (IC50 values) to a series of NTRK inhibitors in MOLM-13 cells, transduced with indicated sgRNAs/Cas9 viruses. D, Drug sensitivity to NTRK inhibitors entrectinib (top) and larotrectinib (bottom) as measured in Fig. 2A and F, in MOLM-13 cells transduced with indicated sgRNAs/Cas9 viruses. E, Expression levels of NTRK1, 2 and 3 mRNA using qRT-PCR analysis with RNAs isolated from MOLM-13 and MV4-11 cells with TP53 KO alleles or nontargeting (NT) control. Expression values were normalized to 36B4 gene-expression levels. Statistical significance was determined by ANOVA with Tukey post-test (**, P < 0.01; ***, P < 0.001). F, Western blot analyses of proteins extracted from MOLM-13 parental and MOLM-13 transduced with TP53 sgRNA/Cas9 viruses and treated overnight with venetoclax (Ven; 100 nmol/L), entrectinib (Entr; 100 nmol/L), or vehicle (DMSO) and identified with antisera to pan-NTRK, phosphorylated NTRK (Tyr516), phosphorylated ERK1/2 (Thr202/Tyr204), ERK1/2, phosphorylated AKT (Thr308), AKT, and GAPDH. G, Western blot analyses of proteins from AML patient samples with known TP53-mutant (TP53mut) or WT (TP53WT) status. Samples 17-00248 and 15-00309 have FLT3-ITD mutations; sample 17-00300 has a FLT3D835E mutation. study using two independent TP53-mutant samples showed data suggest a dependency on new survival pathways (e.g., a response to entrectinib treatment as detected by reduced NTRK) in TP53-mutant cells and patients. Cells resistant to spleen weight and disease burden relative to vehicle control venetoclax exhibited perturbations in mitochondrial homeo­ (Supplementary Fig. S6B). stasis, were abnormally protective in responses to various Together, these data have identified a mechanism of vene­ mitochondrial stressors, and exhibited significant metabolic toclax resistance in AML, namely, the p53 apoptosis machin­ differences, indicative of high proliferation rates, energy pro­ ery, and demonstrated an approach to identify and validate duction, and DNA synthesis. These metabolic differences candidate genes involved in drug resistance. In addition, our were echoed in AML patient samples with TP53 mutations.

OF11 | CANCER DISCOVERY JULY 2019 www.aacrjournals.org

Downloaded from cancerdiscovery.aacrjournals.org on September 29, 2021. © 2019 American Association for Cancer Research. Published OnlineFirst May 2, 2019; DOI: 10.1158/2159-8290.CD-19-0125

TP53 Network Mediates Resistance to BCL2 Inhibition RESEARCH ARTICLE

DISCUSSION detection of the TRK receptors on immature leukemic cell lines demonstrated that NTRK1 expression and activation are not Venetoclax therapy for CLL has produced durable responses restricted to the nervous system, but are also an important in a majority of patients (79%), with complete response in 20% element in early hematopoiesis (63). TRK receptors, which pro­ of patients (60). Venetoclax has shown limited success in AML mote the proliferation and survival of erythroblasts, dendritic clinical trials as a single agent due to drug resistance in the cells, lymphocytes, monocytes, and macrophages (64), have also majority of patients (24). Both of these clinical results are mir­ been detected in human bone marrow cells (65). TRK expres­ rored in venetoclax sensitivity observed ex vivo in patient sam­ sion occurs in myeloid leukemia cell lines and in primary leu­ ples assayed in the Beat AML cohort (40), where we observed kemic cells from patients with AML (66). Subsequent studies greater activity of venetoclax on CLL (141 patients) than on have shown that activation of TRK receptors in hematologic AML (289 patients) specimens, with a median IC for AML sig­ 50 malignancies could result from chromosomal rearrangements nificantly higher than that for CLL (1.4 μmol/L vs. 0.1 μmol/L, (67), point mutations (68), truncations (69), and transcriptional respectively; Mann–Whitney P < 0.001). The efficacy of vene­ changes (70). Although the prevalence of NTRK-mediated can­ toclax in newly diagnosed patients with AML, when adminis­ cers is <1%, fusions involving NTRK3 were observed in various tered in combination with azacytidine, has been more robust, hematologic malignancies including AML (67). The efficacy of although the TP53-mutant subset exhibited the worst response TRK inhibitors has been demonstrated in NTRK fusion–positive rate of any subset of patients reported in the study (25, 61), and human leukemia cell lines and patient-derived xenograft studies, this combination, as well as other venetoclax combinations, is highlighting the potential clinical utility of these inhibitors for a likely to see broader use in AML and in CLL, prompting a need subset of patients with leukemia (71, 72). to better understand possible mechanisms of resistance. Our In our study, sensitivity to TRK inhibitors correlated with an findings identify genes affecting control and execution of apop­ increase in expression of NTRK3 and a concomitant decrease tosis in mitochondria (TP53, BAX, and PMAIP1); their inacti­ in NTRK1 in TP53 KO cells at the RNA and protein lev­ vation individually establishes venetoclax resistance in AML els, indicating a transcriptional regulation. We also observed cell lines. These results correlate with a recent study reporting increased TRK protein levels in TP53-mutant patient samples. activation of p53 through inhibition of MDM2 overcomes Among genes that regulate NTRK expression, RUNX family venetoclax resistance in AML cell lines (30). The mechanism of proteins bind NTRK loci (73) and have regulatory roles in both venetoclax action is concentrated on the mitochondria, a center hematopoietic development and cancer (74). RUNX1 (AML1), of energy production through cellular respiration as well as ini­ a frequently mutated gene in AML, is often disrupted by trans­ tiation of apoptosis; its target, BCL2, functions in both cell sur­ location (75). RUNX1 isoforms can either enhance or repress vival and the oxidative phosphorylation process. In many ways, gene expression in AML cell lines (76) with isoform RUNX1A homeostasis of mitochondria and cell survival in apoptosis are having a repressor function due to the absence of its C-ter­ tightly linked. Attesting to this point, venetoclax-resistant cells minal transactivation domain. AML1 expression is decreased show both: likely resistance through changes in expression of in both TP53-mutant patient samples and TP53 KO cell lines prosurvival and proapoptotic proteins and perturbation in (Supplementary Fig. S6C and S6D), suggesting a potential mitochondrial stress sensitivities. mechanism whereby the interplay between various RUNX iso­ forms may regulate NTRK expression levels. In mice, p53 binds Loss of p53 Affects Sensitivity to Other Signaling to the distal Runx1 promoter, leading to AML1 upregulation Pathway Inhibitors and lymphoma development (77). Future studies will examine Loss of p53 affected drug sensitivities to a wide range of small whether NTRK loci occupancy with various RUNX1 isoforms molecules, including FLT3 inhibitors. Many tyrosine kinase leads to upregulation of specific NTRKs inTP53 -mutant cells. inhibitors have multiple targets, making it difficult to dissect sensitivity change unless a group of related inhibitors is affected. Genome-Wide Screen Identified Genes For example, a previous study of patients with relapsed/refrac­ Controlling Proapoptotic Responses and tory AML with newly acquired crenolanib resistance identified Mitochondria Functions TP53 mutations associated with resistance (62). Our results Our screen identified genes whose inactivation confers vene­ expand on that finding and include loss of sensitivity changes toclax resistance. The most significant hits have a similar mode for many FLT3 inhibitors, an observation that is likely due to of action in apoptosis, namely, a strong proapoptotic activity. disruption of the normal apoptotic response in TP53-mutant These included p53, a transcription factor with both activating cells. Additionally, we observed a new dependency on the NTRK and repressing functions controlling many aspects of cellular pathway, which was confirmed by a similar increase in sensi­ responses, including genotoxic stress pertinent to apoptotic tivity to several inhibitors, either affecting NTRK/ALK/ROS1 signaling, and several p53 transcriptional targets, BAX, BAK1/ signaling such as entrectinib, or the NTRK-selective inhibitor BAK, and PMAIP1 (37, 78, 79). We found changes in the gene larotrectinib. TP53 mutations in AML are typically rare (<8% expression of several p53-regulated genes altered in our TP53 of all cases); however, a dysfunctional WT TP53 is much more KO MOLM-13 cells, including proapoptotic PMAIP1, BAK1, common and requires functional assessment of TP53. In general, and PUMA, without notable changes to BAX expression. BAX TP53 mutations confer poor prognosis and reduced sensitivity and BAK are critical for both mitochondrial outer membrane to a variety of small molecules, and frequently associate with permeabilization (MOMP) in apoptosis and mitochondrial per­ relapsed/refractory AML cases (62). meability transition pore (MPTP) formation. PMAIP1 senses Our future studies will focus on assessing NTRK inhibitor apoptotic signaling and inhibits MCL1. TFDP1, together with sensitivity in TP53-mutant AML patient sample xenografts. The , activates and translocates PMAIP1 into mitochondria

JULY 2019 CANCER DISCOVERY | OF12

Downloaded from cancerdiscovery.aacrjournals.org on September 29, 2021. © 2019 American Association for Cancer Research. Published OnlineFirst May 2, 2019; DOI: 10.1158/2159-8290.CD-19-0125

RESEARCH ARTICLE Nechiporuk et al.

(80), and SLC25A6 (ANT3), a nucleotide ADP/ATP translo­ correlating with decreased levels of PUMA in TP53 KO cells case that is involved in formation of MPTP, is important for and the observed metabolic changes (44). Several other changes depolarization in mitochondrial recycling (refs. 81–83; Fig. 6E). in metabolic profile, namely increases or decreases in several Importantly, venetoclax induced membrane depolarization in amino acids (variable between TP53- and BAX-mutant cells) control cells (Fig. 5) and showed synergy with CCCP in mutant and urea cycle intermediates (common to both TP53- and BAX- cells, suggesting that it is part of its mechanism of action. In mutant cells), suggest the presence of other catabolic shunts to fact, we found that mitochondria in venetoclax-resistant cells provide carbon sources for amino acids necessary for new cell show a different response to mitochondrial stress, and both production. Metabolic perturbations similar to those observed TP53- and BAX-mutant cells had decreased mitophagy and in TP53 KO cells were found in our primary AML patient sam­ exhibited membrane depolarization protection. ples. It has been observed that treatment of leukemic stem cells In TP53 KO cells, we found changes beyond direct p53 target (LSC) from newly diagnosed patients with AML with venetoclax genes. These included downregulation of BCL2 and MCL1 and and a DNMT inhibitor (DNMTi) leads to reduced amino acid significant upregulation of BCLXL, without notable changes levels, inhibition of oxidative phosphorylation, and cell death to BAX expression. With the exception of MCL1, the upregula­ (43). Furthermore, LSCs obtained from patients with AML tion of BCLXL and downregulation of BCL2 correlate well with are resistant to venetoclax and a DNMTi due to their ability expression levels of p53 in a large AML cohort (40) as well as in to compensate for amino acid loss by increasing fatty-acid pediatric acute lymphoblastic leukemia (39). These alterations metabolism (43), an effect we also observed in TP53 KO cells in expression of proapoptotic proteins may reflect compensa­ and in AML patient samples. Given that DNMTi incorporate tory changes. TP53 KO cells were sensitive to the BCL2/BCLXL into DNA and inhibit DNA methylation, it may be that the inhibitor (AZD-4320); however, BAX KO cells were 10-fold less increased nucleotide synthesis observed in TP53 KO cells ele­ sensitive. In contrast, TP53 KO, but not BAX KO, cells were less vates competition between nucleotides and chemical nucleotide sensitive to a selective MCL1 inhibitor (AZD-5991). These dif­ analogues, thereby rendering DNMT inhibitors less effective. ferences may reflect upregulation of BCLXL and a decrease in MCL1 levels in TP53 KO cells but not in BAX KO cells. With Clinically Relevant Hits Are Identified regard to BAX KO insensitivity to venetoclax, although both within the Beat AML Dataset BAX and BAK can result in MOMP, our data may suggest Identification of a significant collection of apoptosis and functional differences between BCL2-driven and MCL1-driven mitochondria homeostasis controlling genes in our genome- BAX and BAK inhibition. The mechanism may involve localiza­ wide screen for venetoclax resistance validated our top hit find­ tion of BAX (cytosolic) and BAK (mitochondrial) prior to the ings. However, although they provide mechanistic insight, some oligomerization processes required for MOMP. validated hits do not correlate with available patient sample data. For example, we did not find a correlation between low Metabolic Changes in Venetoclax-Resistant Cells PMAIP1 expression and low venetoclax sensitivity in our AML Likely Reflect an Increased Proliferation Rate in patient cohort, nor were mutations in PMAIP1 associated with TP53 and BAX KO Cells with Defective Apoptosis loss of venetoclax sensitivity, underlying the importance of lev­ Metabolic profiling of our mutant cells further indicated eraging an independent patient data resource to the screen. changes in energy utilization. Perturbations in the metabo­ Although some hits identified in the screen are not represented lism of cancer cells have been long recognized; for example, as either mutations or low expressers in the Beat AML dataset, some cancer cells rely on glycolysis and pentose phosphate collectively they identify the larger network of genes mechanisti­ pathways versus cellular respiration, known as the Warburg cally responsible for venetoclax sensitivity. Although we focused effect (48). This phenomenon is directed toward utilization of on the loss of function of genes conferring venetoclax resistance, carbon sources for nucleotides, fatty acid and protein synthesis our screen predicts that gain-of-function mutations in prosur­ required for cellular proliferation. Although MOLM-13 cells vival genes in the apoptotic cascade will also result in resistance are leukemia-derived with a cancer metabolism, inactivation of to venetoclax. Notably, a companion study in this issue by Chen p53 and BAX further exacerbated metabolic changes: Increased and colleagues, where the focus of investigation was on a discov­ cellular proliferation led to a diversion of resources into pro­ ery of venetoclax-sensitizing genes, revealed genes in pertinent duction of nucleotides, fatty acids, and proteins required for mitochondrial functions. Further analysis of the venetoclax- membrane production. Curiously, this was accompanied by sensitizing genes in this companion study uncovered impacts increased cellular respiration/increase in ROS production rela­ on mitochondrial function and metabolic pathways that were tive to parental cells, suggesting that mitochondria are over­ opposite to the affects observed with venetoclax-resistance genes compensating for the glycolysis intermediates shunting into here. Collectively, these data suggest that mitochondrial func­ the pentose phosphate pathway. Similar metabolic changes, tion and specific metabolic pathways are essential in governing accompanied by increased oxidative phosphorylation as a result both sensitization and resistance to venetoclax. of p53 inactivation, were previously observed in other tissues Recent clinical trial success with the combination of vene­ (84). It is possible that p53 inactivation also has a transcrip­ toclax and a DNMTi (25) identified lower response rates in tional impact on the observed metabolic changes, because it patients with TP53 mutations than in the overall cohort (47% vs. is known to negatively control glycolysis through the induc­ ∼70% overall), aligning with our findings here. Indeed, patients tion of TIGAR expression (45), which degrades fructose-2,6-bi­ in the TP53 mutations subset exhibited the worst response rate sphosphate and opposes the Warburg effect. Similarly, the of any subset of patients reported in the study with all other TP53–PUMA axis has been implicated in metabolic switching molecular/clinical groups in the study exhibiting 60% to 91% toward suppressing oxidative phosphorylation and glycolysis complete response/complete response with incomplete blood

OF13 | CANCER DISCOVERY JULY 2019 www.aacrjournals.org

Downloaded from cancerdiscovery.aacrjournals.org on September 29, 2021. © 2019 American Association for Cancer Research. Published OnlineFirst May 2, 2019; DOI: 10.1158/2159-8290.CD-19-0125

TP53 Network Mediates Resistance to BCL2 Inhibition RESEARCH ARTICLE count recovery (CR/CRi) rate (including the poor cytogenetic Disclosure of Potential Conflicts of Interest risk group). Further, the TP53-mutant patient group had the A.V. Danilov has received commercial research grants from Gilead lowest odds ratio for CR/CRi response and was the only variable Sciences, Verastem, Takeda, AstraZeneca, and Verastem Oncology, that significantly correlated with worse response. A longer study and is a consultant/advisory board member for AbbVie, AstraZeneca, may elucidate whether those TP53-mutant patients receiving Genentech, Verastem Oncology, Seattle Genetics, TG Therapeutics, the combination treatment will relapse at a higher rate, possibly Curis, Celgene, Teva Oncology, and Gilead Sciences. B.J. Druker with a clone that has higher allele frequency TP53 mutation. serves on the board of directors of Amgen, Burroughs Wellcome Fund, and CureOne; reports receiving other commercial research sup­ port from Novartis, Bristol-Myers Squibb, and Pfizer (institutional METHODS funding—PI or co-investigator on clinical trials funded via contract A complete Methods section is provided in Supplementary Material. with OHSU); has ownership interest (including stock, patents, etc.) in Amgen, Blueprint Medicines, MolecularMD (inactive—acquired Cell Lines by ICON Laboratories), GRAIL, Patent 6958335 (exclusively licensed Human MOLM-13 cells were obtained from the Sanger Institute to Novartis), Henry Stewart Talks, Merck via Dana-Farber Cancer Cancer Cell Line Panel. Human MV4-11, OCI-AML2, and THP-1 Institute (royalty payments); is a consultant/advisory board member were obtained from ATCC. Authentication was performed on all cell for Aileron Therapeutics, ALLCRON, Third Coast Therapeutics, Vivid lines used in this study at the Oregon Health and Science University Biosciences, Monojul (inactive), Baxalta (inactive), CTI Biopharma (OHSU) DNA Services Core facility. All cell lines were tested for Myco- (inactive), Aptose, Beta Cat, Blueprint Medicines, Celgene, Cepheid, plasma on a monthly schedule. GRAIL (former), Gilead (former), and Patient True Talk; and is an uncompensated joint steering committee member for Beat AML CRISPR/Cas9 Screen LLC. J.W. Tyner has received commercial research grants from Agios, Loss-of-function screens were performed using two human Aptose, Array, AstraZeneca, Constellation, Genentech, Gilead, Incyte, genome-wide sgRNA libraries—Y. Kosuke library (31) and Brunello Janssen, Seattle Genetics, Syros, and Takeda; has received honoraria library (33)—both purchased form Addgene (#73179 and #67989, from the speakers bureaus of Therapeutic Advances in Childhood respectively). Leukemia and Hermedicus - Acute Leukemia Forum; has ownership interest in Vivid Biosciences; and is a consultant/advisory board Biostatistical Analysis member of Vivid Biosciences. No potential conflicts of interest were disclosed by the other authors. Pipeline for executing analyses of CRISPR library sequences was performed using MAGeCK analyses (32). Authors’ Contributions Ex Vivo Functional Screen Conception and design: T. Nechiporuk, S.E. Kurtz, C.E. Tognon, Small-molecule inhibitors, purchased from LC Laboratories Inc. and J.W. Tyner Development of methodology: T. Nechiporuk, T. Liu, A. D’Alessandro, Selleck Chemicals, were reconstituted in DMSO and stored at −80°C. S.K. Joshi Xenograft Generation Acquisition of data (provided animals, acquired and managed patients, provided facilities, etc.): T. Nechiporuk, S.E. Kurtz, T. Liu, C.L. Animal studies were conducted with approval from the OHSU Jones, A. D’Alessandro, A. d’Almeida, S.K. Joshi, A.V. Danilov, B.H. Chang Institutional Animal Care and Use Committee. Patient samples were Analysis and interpretation of data (e.g., statistical analysis, injected into cohorts of NSGS mice and allowed to engraft until the biostatistics, computational analysis): T. Nechiporuk, S.E. Kurtz, peripheral blood showed approximately 1% human CD33 chimerisms. O. Nikolova, T. Liu, A. D’Alessandro, R. Culp-Hill, S.K. Joshi, M. Metabolomic Analysis Rosenberg, A.V. Danilov, J.W. Tyner Writing, review, and/or revision of the manuscript: T. Nechiporuk, Two million MOLM-13 and MV4-11 cells were flash-frozen as cell S.E. Kurtz, O. Nikolova, T. Liu, R. Culp-Hill, S.K. Joshi, C.E. Tognon, pellet and metabolomics analyses were performed via ultra-high- B.J. Druker, B.H. Chang, J.W. Tyner pressure liquid chromatography–mass spectrometry (UHPLC-MS; Administrative, technical, or material support (i.e., reporting or Vanquish and Q Exactive, Thermo Fisher) as previously reported (85). organizing data, constructing databases): T. Nechiporuk, R. Culp-Hill, J.W. Tyner Mitophagy Detection Study supervision: A. D’Alessandro, B.J. Druker, J.W. Tyner Cells were loaded in complete medium with 100 nmol/L Mitophagy Other (development of prioritization framework to assist in Dye (Mitophagy Detection Kit, Dojindo Molecular Technologies). rigor and reproducibility): S.K. McWeeney The data were analyzed post acquisition using FlowJo software (Tree Star). Ten thousand cells were acquired using channels 405-2-A (lyso Acknowledgments Dye), 561-3-A (Mitophagy Dye), and 640-1-A (MitoTracker Dye). We thank Wesley Horton and Kevin Watanabe-Smith for analytical Mitochondrial Membrane Potential support and the OHSU Massively Parallel Sequencing Shared Resource and Flow Cytometry Core for technical support. This study was also Mitochondrial membrane potential was assessed using tetrameth­ supported by grants from the NCI (1U01CA217862, 1U54CA224019, ylrhodamine methyl ester bound to cationic carbocyanine (JC-1, and 3P30CA069533-18S5). J.W. Tyner received grants from the V Foun­ Dojingo Molecular Technology). Cells were treated with 100 nmol/L dation for Cancer Research, the Gabrielle’s Angel Foundation for Can­ venetoclax or DMSO and with or without CCCP for 2 hours prior to cer Research, and the NCI (1R01CA183947). S.K. Joshi is supported by the addition of the dye and flow cytometry. the ARCS Scholar Foundation and the NCI (1F30CA239335-10).

Data Availability The costs of publication of this article were defrayed in part by Raw data files for CRISPR screens have been deposited at the Gene the payment of page charges. This article must therefore be hereby Expression Omnibus and can be found under the accession number marked advertisement in accordance with 18 U.S.C. Section 1734 GSE130414. solely to indicate this fact.

JULY 2019 CANCER DISCOVERY | OF14

Downloaded from cancerdiscovery.aacrjournals.org on September 29, 2021. © 2019 American Association for Cancer Research. Published OnlineFirst May 2, 2019; DOI: 10.1158/2159-8290.CD-19-0125

RESEARCH ARTICLE Nechiporuk et al.

Received January 30, 2019; revised April 20, 2019; accepted April 22. Roberts AW, Davids MS, Pagel JM, Kahl BS, Puvvada SD, Gerecitano 30, 2019; published first May 2, 2019. JF, et al. Targeting BCL2 with venetoclax in relapsed chronic lympho­ cytic leukemia. N Engl J Med 2016;374:311–22. 23. Seymour JF, Ma S, Brander DM, Choi MY, Barrientos J, Davids MS, et al. Venetoclax plus in relapsed or refractory chronic REFERENCES lymphocytic leukaemia: a phase 1b study. Lancet Oncol 2017;18:230– . 1 Hanahan D, Weinberg RA. Hallmarks of cancer: the next generation. 240. Cell 2011;144:646–74. 24. Konopleva M, Pollyea DA, Potluri J, Chyla B, Hogdal L, Busman T, 2. Llambi F, Green DR. Apoptosis and oncogenesis: give and take in the et al. Efficacy and biological correlates of response in a phase II study BCL-2 family. Curr Opin Genet Dev 2011;21:12–20. of venetoclax monotherapy in patients with acute myelogenous leu­ 3. Hata AN, Engelman JA, Faber AC. The BCL2 family: key mediators kemia. Cancer Discov 2016;6:1106–1117. of the apoptotic response to targeted anticancer therapeutics. Cancer 25. DiNardo CD, Pratz KW, Letai A, Jonas BA, Wei AH, Thirman M, et al. Discov 2015;5:475–87. Safety and preliminary efficacy of venetoclax with decitabine or azac­ 4. Letai AG. Diagnosing and exploiting cancer’s addiction to blocks in itidine in elderly patients with previously untreated acute myeloid apoptosis. Nat Rev Cancer 2008;8:121–32. leukaemia: a non-randomised, open-label, phase 1b study. Lancet 5. Khoo KH, Verma CS, Lane DP. Drugging the p53 pathway: under­ Oncol 2018;19:216–228. standing the route to clinical efficacy. Nat Rev Drug Discov 2014; 26. Holohan C, Van Schaeybroeck S, Longley DB, Johnston PG. Cancer 13:217–36. drug resistance: an evolving paradigm. Nat Rev Cancer 2013;13:714–26. 6. Haupt Y, Maya R, Kazaz A, Oren M. Mdm2 promotes the rapid deg­ 27. Phillips DC, Xiao Y, Lam LT, Litvinovich E, Roberts-Rapp L, Souers radation of p53. Nature 1997;387:296–9. AJ, et al. Loss in MCL-1 function sensitizes non-Hodgkin’s lym­ 7. Wertz IE, Kusam S, Lam C, Okamoto T, Sandoval W, Anderson DJ, phoma cell lines to the BCL-2-selective inhibitor venetoclax (ABT- et al. Sensitivity to antitubulin chemotherapeutics is regulated by 199). Blood Cancer J 2015;5:e368. MCL1 and FBW7. Nature 2011;471:110–4. 28. Bodo J, Zhao X, Durkin L, Souers AJ, Phillips DC, Smith MR, et al. 8. Rochaix P, Krajewski S, Reed JC, Bonnet F, Voigt JJ, Brousset P. In vivo Acquired resistance to venetoclax (ABT-199) in t(14;18) positive lym­ patterns of Bcl-2 family protein expression in breast carcinomas in phoma cells. Oncotarget 2016;7:70000–70010. relation to apoptosis. J Pathol 1999;187:410–5. 29. Lin KH, Winter PS, Xie A, Roth C, Martz CA, Stein EM, et al. Target­ 9. Zhang B, Gojo I, Fenton RG. Myeloid cell factor-1 is a critical survival ing MCL-1/BCL-XL forestalls the acquisition of resistance to ABT- factor for multiple myeloma. Blood 2002;99:1885–93. 199 in acute myeloid leukemia. Sci Rep 2016;6:27696. 10. Beroukhim R, Mermel CH, Porter D, Wei G, Raychaudhuri S, Donovan J, 30. Pan R, Ruvolo V, Mu H, Leverson JD, Nichols G, Reed JC, et al. Syn­ et al. The landscape of somatic copy-number alteration across human thetic lethality of combined Bcl-2 inhibition and p53 activation in cancers. Nature 2010;463:899–905. AML: mechanisms and superior antileukemic efficacy. Cancer Cell 11. Kojima K, Konopleva M, McQueen T, O’Brien S, Plunkett W, Andreeff 2017;32:748–760. M. Mdm2 inhibitor Nutlin-3a induces p53-mediated apoptosis by 31. Tzelepis K, Koike-Yusa H, De Braekeleer E, Li Y, Metzakopian E, transcription-dependent and transcription-independent mechanisms Dovey OM, et al. A CRISPR dropout screen identifies genetic vulner­ and may overcome Atm-mediated resistance to fludarabine in chronic abilities and therapeutic targets in acute myeloid leukemia. Cell Rep lymphocytic leukemia. Blood 2006;108:993–1000. 2016;17:1193–1205. 12. Hasegawa H, Yamada Y, Iha H, Tsukasaki K, Nagai K, Atogami S, et al. 32. Li W, Xu H, Xiao T, Cong L, Love MI, Zhang F, et al. MAGeCK enables Activation of p53 by Nutlin-3a, an antagonist of MDM2, induces robust identification of essential genes from genome-scale CRISPR/ apoptosis and cellular senescence in adult T-cell leukemia cells. Leu­ Cas9 knockout screens. Genome Biol 2014;15:554. kemia 2009;23:2090–101. 33. Doench JG, Fusi N, Sullender M, Hegde M, Vaimberg EW, Donovan 13. Tait SW, Green DR. Mitochondria and cell death: outer membrane KF, et al. Optimized sgRNA design to maximize activity and mini­ permeabilization and beyond. Nat Rev Mol Cell Biol 2010;11:621–32. mize off-target effects of CRISPR-Cas9. Nat Biotechnol 2016;34: 14. Juin P, Geneste O, Gautier F, Depil S, Campone M. Decoding and 184–191. unlocking the BCL-2 dependency of cancer cells. Nat Rev Cancer 34. Woo IS, Jin H, Kang ES, Kim HJ, Lee JH, Chang KC, et al. TMEM14A 2013;13:455–65. inhibits N-(4-hydroxyphenyl)retinamide-induced apoptosis through 15. Oltersdorf T, Elmore SW, Shoemaker AR, Armstrong RC, Augeri DJ, the stabilization of mitochondrial membrane potential. Cancer Lett Belli BA, et al. An inhibitor of Bcl-2 family proteins induces regression 2011;309:190–8. of solid tumours. Nature 2005;435:677–81. 35. Hershko T, Ginsberg D. Up-regulation of Bcl-2 homology 3 (BH3)-only 16. Souers AJ, Leverson JD, Boghaert ER, Ackler SL, Catron ND, Chen J, proteins by E2F1 mediates apoptosis. J Biol Chem 2004;279:8627–34. et al. ABT-199, a potent and selective BCL-2 inhibitor, achieves anti­ 36. Imazu T, Shimizu S, Tagami S, Matsushima M, Nakamura Y, Miki T, tumor activity while sparing platelets. Nat Med 2013;19:202–8. et al. Bcl-2/E1B 19 kDa-interacting protein 3-like protein (Bnip3L) 17. Caenepeel S, Brown SP, Belmontes B, Moody G, Keegan KS, Chui D, interacts with bcl-2/Bcl-xL and induces apoptosis by altering mito­ et al. AMG 176, a Selective MCL1 inhibitor, is effective in hematologic chondrial membrane permeability. Oncogene 1999;18:4523–9. cancer models alone and in combination with established therapies. 37. Fischer MCensus and evaluation of p53 target genes. Oncogene 2017; Cancer Discov 2018;8:1582–1597. 36:3943–3956. 18. Kotschy A, Szlavik Z, Murray J, Davidson J, Maragno AL, Le Toumelin- 38. Kojima K, Konopleva M, Samudio IJ, Shikami M, Cabreira-Hansen M, Braizat G, et al. The MCL1 inhibitor S63845 is tolerable and effective McQueen T, et al. MDM2 antagonists induce p53-dependent apoptosis in diverse cancer models. Nature 2016;538:477–482. in AML: implications for leukemia therapy. Blood 2005;106:3150–9. 19. Ramsey HE, Fischer MA, Lee T, Gorska AE, Arrate MP, Fuller L, et al. 39. Findley HW, Gu L, Yeager AM, Zhou M. Expression and regulation A Novel MCL1 inhibitor combined with venetoclax rescues veneto­ of Bcl-2, Bcl-xl, and Bax correlate with p53 status and sensitivity to clax-resistant acute myelogenous leukemia. Cancer Discov 2018;8: apoptosis in childhood acute lymphoblastic leukemia. Blood 1997;89: 1566–1581. 2986–93. 20. Tron AE, Belmonte MA, Adam A, Aquila BM, Boise LH, Chiarparin E, 40. Tyner JW, Tognon CE, Bottomly D, Wilmot B, Kurtz SE, Savage SL, et al. Discovery of Mcl-1-specific inhibitor AZD5991 and preclinical et al. Functional genomic landscape of acute myeloid leukaemia. activity in multiple myeloma and acute myeloid leukemia. Nat Com­ Nature 2018;562:526–531. mun 2018;9:5341. 41. Blankenberg FG. In vivo detection of apoptosis. J Nucl Med 2008; 21. Pan R, Hogdal LJ, Benito JM, Bucci D, Han L, Borthakur G, et al. 49:81s–95s. Selective BCL-2 inhibition by ABT-199 causes on-target cell death in 42. Lagadinou ED, Sach A, Callahan K, Rossi RM, Neering SJ, Minhajud­ acute myeloid leukemia. Cancer Discov 2014;4:362–75. din M, et al. BCL-2 inhibition targets oxidative phosphorylation and

OF15 | CANCER DISCOVERY JULY 2019 www.aacrjournals.org

Downloaded from cancerdiscovery.aacrjournals.org on September 29, 2021. © 2019 American Association for Cancer Research. Published OnlineFirst May 2, 2019; DOI: 10.1158/2159-8290.CD-19-0125

TP53 Network Mediates Resistance to BCL2 Inhibition RESEARCH ARTICLE

selectively eradicates quiescent human leukemia stem cells. Cell Stem product/NGF receptor in undifferentiated hematopoietic cells. Blood Cell 2013;12:329–41. 1994;83:1479–85. 43. Jones CL, Stevens BM, D’Alessandro A, Reisz JA, Culp-Hill R, Nem­ 64. Vega JA, García-Suárez O, Hannestad J, Pérez-Pérez M, Germanà A. kov T, et al. Inhibition of amino acid metabolism selectively targets Neurotrophins and the immune system. J Anat 2003;203:1–19. human leukemia stem cells. Cancer Cell 2018;34:724–740. 65. Labouyrie E, Dubus P, Groppi A, Mahon FX, Ferrer J, Parrens M, et al. 44. Kim J, Yu L, Chen W, Xu Y, Wu M, Todorova D, et al. Wild-type p53 Expression of neurotrophins and their receptors in human bone mar­ promotes cancer metabolic switch by inducing PUMA-dependent sup­ row. Am J Pathol 1999;154:405–15. pression of oxidative phosphorylation. Cancer Cell 2019;35:191–203. 66. Kaebisch A, Brokt S, Seay U, Lohmeyer J, Jaeger U, Pralle H. Expres­ 45. Bensaad K, Tsuruta A, Selak MA, Vidal MN, Nakano K, Bartrons R, sion of the nerve growth factor receptor c-TRK in human myeloid et al. TIGAR, a p53-inducible regulator of glycolysis and apoptosis. leukaemia cells. Br J Haematol 1996;95:102–9. Cell 2006;126:107–20. 67. Eguchi M, Eguchi-Ishimae M, Tojo A, Morishita K, Suzuki K, 46. Polyak K, Xia Y, Zweier JL, Kinzler KW, Vogelstein B. A model for p53- Sato Y, et al. Fusion of ETV6 to neurotrophin-3 receptor TRKC induced apoptosis. Nature 1997;389:300–5. in acute myeloid leukemia with t(12;15)(p13;q25). Blood 1999;93: 47. Li PF, Dietz R, von Harsdorf R. p53 regulates mitochondrial membrane 1355–63. potential through reactive oxygen species and induces cytochrome 68. Tomasson MH, Xiang Z, Walgren R, Zhao Y, Kasai Y, Miner T, et al. c-independent apoptosis blocked by Bcl-2. Embo J 1999;18: Somatic mutations and germline sequence variants in the expressed 6027–36. tyrosine kinase genes of patients with de novo acute myeloid leuke­ 48. Vander Heiden MG, Cantley LC, Thompson CB. Understanding the mia. Blood 2008;111:4797–808. Warburg effect: the metabolic requirements of cell proliferation. Sci­ 69. Reuther GW, Lambert QT, Caligiuri MA, Der CJ. Identification and ence 2009;324:1029–33. characterization of an activating TrkA deletion mutation in acute 49. Kirshner JR, He S, Balasubramanyam V, Kepros J, Yang CY, Zhang myeloid leukemia. Mol Cell Biol 2000;20:8655–66. M, et al. Elesclomol induces cancer cell apoptosis through oxidative 70. Li Z, Beutel G, Rhein M, Meyer J, Koenecke C, Neumann T, et al. stress. Mol Cancer Ther 2008;7:2319–27. High-affinity neurotrophin receptors and ligands promote leukemo­ 50. Ardini E, Menichincheri M, Banfi P, Bosotti R, De Ponti C, Pulci R, genesis. Blood 2009;113:2028–37. et al. Entrectinib, a pan-TRK, ROS1, ALK inhibitor with activity in 71. Smith KM, Fagan PC, Pomari E, Germano G, Frasson C, Walsh C, multiple molecularly defined cancer indications. Mol Cancer Ther et al. Antitumor activity of entrectinib, a pan-TRK, ROS1, ALK inhibi­ 2016;15:628–39. tor, in ETV6-NTRK3-positive acute myeloid leukemia. Mol Cancer 51. Vaishnavi A, Capelletti M, Le AT, Kako S, Butaney M, Ercan D, et al. Ther 2018;17:455–463. Oncogenic and drug-sensitive NTRK1 rearrangements in lung cancer. 72. Taylor J, Pavlick D, Yoshimi A, Marcelus C, Chung SS, Hechtman JF, Nat Med 2013;19:1469–1472. et al. Oncogenic TRK fusions are amenable to inhibition in hemato­ 52. Taipale M, Krykbaeva I, Whitesell L, Santagata S, Zhang J, Liu Q, et al. logic malignancies. J Clin Invest 2018;128:3819–3825. Chaperones as thermodynamic sensors of drug-target interactions 73. Yevshin I, Sharipov R, Kolmykov S, Kondrakhin Y, Kolpakov F. reveal kinase inhibitor specificities in living cells. Nat Biotechnol GTRD: a database on gene transcription regulation-2019 update. 2013;31:630–7. Nucleic Acids Res 2019;47:D100–d105. 53. Karaman MW, Herrgard S, Treiber DK, Gallant P, Atteridge CE, 74. Wang Q, Stacy T, Binder M, Marin-Padilla M, Sharpe AH, Speck NA. Campbell BT, et al. A quantitative analysis of kinase inhibitor selectiv­ Disruption of the Cbfa2 gene causes necrosis and hemorrhaging in ity. Nat Biotechnol 2008;26:127–32. the central nervous system and blocks definitive hematopoiesis. Proc 54. Hedvat M, Huszar D, Herrmann A, Gozgit JM, Schroeder A, Sheehy Natl Acad Sci U S A 1996;93:3444–9. A, et al. The JAK2 inhibitor AZD1480 potently blocks Stat3 signaling 75. TCGA, genomic and epigenomic landscapes of adult de novo acute and oncogenesis in solid tumors. Cancer Cell 2009;16:487–97. myeloid leukemia. N Engl J Med 2013;368:2059–74. 55. Tognon C, Garnett M, Kenward E, Kay R, Morrison K, Sorensen PH. 76. Gattenloehner S, Chuvpilo S, Langebrake C, Reinhardt D, Müller- The chimeric protein tyrosine kinase ETV6-NTRK3 requires both Hermelink HK, Serfling E, et al. Novel RUNX1 isoforms determine Ras-Erk1/2 and PI3-kinase-Akt signaling for fibroblast transforma­ the fate of acute myeloid leukemia cells by controlling CD56 expres­ tion. Cancer Res 2001;61:8909–16. sion. Blood 2007;110:2027–33. 56. Lannon CL, Martin MJ, Tognon CE, Jin W, Kim SJ, Sorensen PH. A 77. Shimizu K, Yamagata K, Kurokawa M, Mizutani S, Tsunematsu Y, highly conserved NTRK3 C-terminal sequence in the ETV6-NTRK3 Kitabayashi I. Roles of AML1/RUNX1 in T-cell malignancy induced oncoprotein binds the phosphotyrosine binding domain of insulin by loss of p53. Cancer Sci 2013;104:1033–8. receptor substrate-1: an essential interaction for transformation. J 78. Vogelstein B, Lane D, Levine AJ. Surfing the p53 network. Nature Biol Chem 2004;279:6225–34. 2000;408:307–10. 57. Davare MA, Tognon CE. Detecting and targeting oncogenic fusion 79. Levine AJ, Oren M. The first 30 earsy of p53: growing ever more com­ proteins in the genomic era. Biol Cell 2015;107:111–29. plex. Nat Rev Cancer 2009;9:749–58. 58. Vaishnavi A, Le AT, Doebele RC. TRKing down an old oncogene in a 80. Joshi-Tope G, Gillespie M, Vastrik I, D’Eustachio P, Schmidt E, de new era of . Cancer Discov 2015;5:25–34. Bono B, et al. Reactome: a knowledgebase of biological pathways. 59. Nakagawara A, Arima-Nakagawara M, Scavarda NJ, Azar CG, Cantor Nucleic Acids Res 2005;33(Database issue):D428–32. AB, Brodeur GM. Association between high levels of expression of the 81. Palmieri F. The mitochondrial transporter family SLC25: identification, TRK gene and favorable outcome in human neuroblastoma. N Engl J properties and physiopathology. Mol Aspects Med 2013;34:465–84. Med 1993;328:847–54. 82. Verrier F, Mignotte B, Jan G, Brenner C. Study of PTPC composition 60. Edelmann J, Gribben JG. Managing patients with TP53-deficient during apoptosis for identification of viral protein target. Ann N Y chronic lymphocytic leukemia. J Oncol Pract 2017;13:371–377. Acad Sci 2003;1010:126–42. 61. DiNardo CD, Pratz K, Pullarkat V, Jonas BA, Arellano M, Becker 83. Rodriguez-Enriquez S, He L, Lemasters JJ. Role of mitochondrial per­ PS, et al. Venetoclax combined with decitabine or azacitidine in meability transition pores in mitochondrial autophagy. Int J Biochem treatment-naive, elderly patients with acute myeloid leukemia. Blood Cell Biol 2004;36:2463–72. 2019;133:7–17. 84. Prokesch A, Graef FA, Madl T, Kahlhofer J, Heidenreich S, Schumann 62. Zhang H, Savage S, Schultz AR, Bottomly D, White L, Segerdell E, A, et al. Liver p53 is stabilized upon starvation and required for amino et al. Clinical resistance to crenolanib in acute myeloid leukemia due acid catabolism and gluconeogenesis. Faseb j 2017;31:732–742. to diverse molecular mechanisms. Nat Commun 2019;10:244. 85. Nemkov T, D’Alessandro A, Hansen KC. Three-minute method for 63. Chevalier S, Praloran V, Smith C, MacGrogan D, Ip NY, Yancopoulos amino acid analysis by UHPLC and high-resolution quadrupole GD, et al. Expression and functionality of the trkA proto-oncogene orbitrap mass spectrometry. Amino Acids 2015;47:2345–57.

JULY 2019 CANCER DISCOVERY | OF16

Downloaded from cancerdiscovery.aacrjournals.org on September 29, 2021. © 2019 American Association for Cancer Research. Published OnlineFirst May 2, 2019; DOI: 10.1158/2159-8290.CD-19-0125

The TP53 Apoptotic Network Is a Primary Mediator of Resistance to BCL2 Inhibition in AML Cells

Tamilla Nechiporuk, Stephen E. Kurtz, Olga Nikolova, et al.

Cancer Discov Published OnlineFirst May 2, 2019.

Updated version Access the most recent version of this article at: doi:10.1158/2159-8290.CD-19-0125

Supplementary Access the most recent supplemental material at: Material http://cancerdiscovery.aacrjournals.org/content/suppl/2019/05/02/2159-8290.CD-19-0125.DC1

E-mail alerts Sign up to receive free email-alerts related to this article or journal.

Reprints and To order reprints of this article or to subscribe to the journal, contact the AACR Publications Subscriptions Department at [email protected].

Permissions To request permission to re-use all or part of this article, use this link http://cancerdiscovery.aacrjournals.org/content/early/2019/06/18/2159-8290.CD-19-0125. Click on "Request Permissions" which will take you to the Copyright Clearance Center's (CCC) Rightslink site.

Downloaded from cancerdiscovery.aacrjournals.org on September 29, 2021. © 2019 American Association for Cancer Research.