Acute leukemia is a rapidly progressing blood cancer with low survival rates. Unfavorable prognosis is attributed to insufficiently characterized subpopulations of leukemia stem cells (LSC) that drive chemoresistance and leukemia relapse. Here we utilized a genetic reporter that assesses stemness to enrich and functionally characterize LSCs. We observed heterogeneous activity of the ERG+85 enhancer–based fluorescent reporter in human leukemias. Cells with high reporter activity (tagBFPHigh) exhibited elevated expression of stemness and chemoresistance genes and demonstrated increased clonogenicity and resistance to chemo- and radiotherapy as compared with their tagBFPNeg counterparts. The tagBFPHigh fraction was capable of regenerating the original cellular heterogeneity and demonstrated increased invasive ability. Moreover, the tagBFPHigh fraction was enriched for leukemia-initiating cells in a xenograft assay. We identified the ubiquitin hydrolase USP9X as a novel ERG transcriptional target that sustains ERG+85–positive cells by controlling ERG ubiquitination. Therapeutic targeting of USP9X led to preferential inhibition of the ERG-dependent leukemias. Collectively, these results characterize human leukemia cell functional heterogeneity and suggest that targeting ERG via USP9X inhibition may be a potential treatment strategy in patients with leukemia.

Significance:

This study couples a novel experimental tool with state-of-the-art approaches to delineate molecular mechanisms underlying stem cell-related characteristics in leukemia cells.

Acute leukemia is a highly aggressive group of blood malignancies that originate from hematopoietic stem cells (HSC). Accumulation of blast cells in the bone marrow due to deregulation of molecular pathways controlling self-renewal and differentiation of immature blood cells is the main feature of leukemia (1, 2). Well-recognized and prognostic genetic heterogeneity, as well as functional variability, exists among the subsets of leukemia cells obtained from the same patient (3, 4). Functional heterogeneity model posits that a fraction of acute leukemia cells displays sufficient regenerative capacity to propagate the disease, withstand chemotherapy, and cause leukemia relapse. Functional resemblance of these leukemic cells to normal hematopoietic stem cells (HSC) contributed to their nomination as leukemia stem cells (LSC; refs. 5, 6).

Although recent studies have shown that phenotypic and genetic heterogeneity within tumors constitutes a major source of therapeutic resistance (7), efficient tools to identify and pull out functional stem cell from the heterogeneous cell population are still lacking. Experimentally, the presence of functional human LSCs can be proved by their capacity to engraft immunodeficient mice and induce leukemia growth in their hematopoietic organs (8, 9). Up to date, enrichment for LSCs has been reached by focusing on: cell surface markers (10), metabolism (11, 12), cell-cycle quiescence (13), and miRNA bioactivity (14). These studies and others demonstrated extraordinary inter- and even intra-sample (3) heterogeneity for LSC activity that was influenced by disease stage and type of therapy (10, 15, 16). Furthermore, prior studies emphasized the need to devise approaches that enable identification and isolation of viable human LSCs based on stemness state of the single cell to pinpoint molecular regulators that maintain LSC properties.

Here, we capitalized on our recent findings that a +85 enhancer of ERG transcription factor (TF) can be used as a probe of cellular stemness state in human normal and leukemic cells when integrated into a lentiviral reporter system (17). Indeed, we and others demonstrated that the endogenous ERG+85 enhancer is particularly active in human HSCs and a subset of leukemias as well as sensitive to the net activity of the multiple TFs, termed heptad, implicated in stemness program regulation (18–20).

Herein, by barcoding leukemia cell lines and patient sample with ERG+85 reporter, we attempted to enrich and characterize leukemia subpopulations endowed with LSC properties that included superior leukemia initiation, invasion, and drug resistance. Gene expression analysis of the subpopulations with different levels of ERG+85 reporter activity uncovered ERG/USP9X feed-forward–regulatory relationships that can be targeted therapeutically.

Full list of references can be found in the Supplementary Data.

Acute myeloid lymphoma patient samples

The study was approved by the Institutional Review Boards of Tel Aviv University (Tel Aviv, Israel) and University Health Network (Toronto, Ontario, Canada). Written informed consent (according to the Declaration of Helsinki) was obtained from all patients. Acute myeloid lymphoma (AML) samples were cultured in StemSpanTM SFEM II medium (StemCell Technologies) supplemented with growth factors [IL3 (10 ng/mL), IL6 (10 ng/mL), G-CSF (10 ng/mL), TPO (25 ng/mL), SCF (50 ng/mL), and FLT3L (50 ng/mL)] on preestablished confluent MS-5 stromal cells.

Cell lines and drug treatments

ELF-153, KASUMI-1, AML193, and ME1 were cultured as recommended by the manufacturer (DSMZ). Jurkat, THP1, and K562 were described elsewhere and grown in RPMI supplemented with FBS (10%), l-glutamine (1%), and penicillin/streptomycin (1%). TEX cell line was grown as described elsewhere (21). All cell lines were authenticated by short tandem repeat profiling using PowerPlex16 HS kit (Promega). Cell number and viability was estimated using hemocytometer and Trypan blue exclusion assay, respectively. For radiation treatment experiments, cells were exposed to ionizing radiation using Cs-137 source at the dose rate 2.8 Gy/minute using GMBH BioBeam 8000 gamma irradiation device (Gamma service). To quantitate clonogenic growth potential of the leukemic cells, we plated Jurkat cells (500 cells/plate for nontreated and 10,000 cells/plate for irradiated) in MethoCult H4100 (StemCell Technologies) supplemented with FBS (30%, MultiCell), 5637 cells’ conditioned medium (10%), l-glutamine (1%), and penicillin/streptomycin (1%), 2-mercaptoethanol (50 μmol/L). Serial replating experiments were performed by harvesting all the colonies from methylcellulose, followed by plating 1,000 cells into new methylcellulose. Colonies were counted under the microscope after 12–14 days of incubation. All cells were maintained in a humidified incubator at 37°C and 5% CO2. All cell lines were tested negative for Mycoplasma by PCR at least once during the period of this study. Ara-C and doxorubicin were purchased from Sigma, WP1130 from Cayman Chemicals, S63845 from Apex, and G9 was provided by Dr. N. Donato (University of Michigan, Ann Arbor, MI).

Leukemia xenotransplantation model

All animal experimental protocols were approved by the Institutional Animal Care and Use Committee of Tel-Aviv University (Tel Aviv, Israel). Jurkat cells were transplanted intrafemorally (IF) as described previously (22) into 8- to 10-week-old NOD-scid IL2Rgnull-3/GM/SF (NSG-SGM3) mice, which were injected intraperitoneally with busulfan (30 mg/kg) 24 hours before transplantation. Human engraftment in the injected and noninjected bones was analyzed 6–8 weeks posttransplantation by flow cytometry analysis using human-specific CD45-Alexa Fluor 750–conjugated antibody, EGFP, and tagBFP positivity. The frequency of repopulating cells was calculated using ELDA software (23).

DNA constructs and cloning

Dual promoter lentiviral reporter vectors (pMIN and pMIN-ERG+85) were described elsewhere (17). The ERG+85 fragment was amplified using Prime GXL high fidelity polymerase (Clontech) from PGL2-ERG+85 vector (24) and cloned into pMIN upstream of mCMV. Full-length human ERG cDNA was subcloned instead of EGFP into a bidirectional lentiviral vector MA1 (25).

USP9X shRNAs were cloned into pLKO1-puro TRC plasmid using EcoRI and AgeI restriction sites. Target sequences for shRNA experiments were CCTAAGGTTAAGTCGCCCTCG (shScramble), CCACCTCAAACCAAGGATCAA (shUSP9X#1), CGCCTGATTCTTCCAATGAAA (shUSP9X #2), and GAGAGTTTATTCACTGTCTTA (shUSP9X #3).

Virus preparation and transduction procedure

Viral particles were generated by transient transfection of 293T cells using CMVDeltaR8.91 and pMD2G constructs as described elsewhere (26). Leukemia cells were infected by addition of viral supernatant to obtain 10%–30% infection rate. Jurkat cells were selected with puromycin (3 μg/mL) for 3 days in the USP9X knockdown experiments. After selection completion, cells were replated for the downstream experiments.

Quantitative RT-PCR and microarrays

RNA was extracted with the TRIzol reagent (Invitrogen) and reverse transcribed with SuperScript III (Invitrogen). Real-time PCR reactions were prepared with SYBR Green PCR Master Mix (Applied Biosystems) in triplicates and analyzed on Applied Biosystems 7900HT instruments. Absolute gene expression was quantified with SDS software (Applied Biosystems) based on the standard curve method and presented after normalization for GAPDH. List of primers is presented in Supplementary Table S1.

RNA from 100,000 or more sorted cells was used for microarray analysis by Human Gene Clariom S Assay (Thermo Fisher Scientific). Data were processed, normalized, and log2 transformed. Microarray raw data are included as Supplementary Data. The data can be viewed in the following link (https://www.ncbi.nlm.nih.gov/geo/query/acc.cgi?acc=GSE131079).

Flow cytometry sorting and analysis

EGFP and tagBFP expression were analyzed using Gallios Flow Cytometer and Cytoflex Flow Cytometer (Beckman Coulter, Inc.) For the intracellular ERG protein-level analysis, cells were fixed in formaldehyde (1.4%), permeabilized in ethanol (100%), followed by labeling with anti-ERG rabbit mAb conjugated with PE (clone A7L1G, Cell Signaling Technology, 1:50) or rabbit anti-IgG control (sc-2027, Santa Cruz Biotechnology). Jurkat cells were sorted using BD FACSAria Fusion sorter (BD Biosciences). In ERG overexpression experiments, anti-CD271-APC (Miltenyi Biotec, 1:50) was used for sorting of ERG-transduced Jurkat cells.

Apoptosis analysis

Apoptosis was measured using Annexin V (Invitrogen) and Zombie NIR Fixable Viability Dye (BioLegend) according to the manufacturer's protocols. Cells were analyzed with Gallios flow cytometer (Beckman Coulter). Flow cytometry data were analyzed using Kaluza flow cytometry analysis software (Beckman Coulter).

Western blotting analysis

The following primary antibodies were used: mouse anti-β-actin (clone 8H10D10, Cell Signaling Technology, 1:2,000), rabbit anti-USP9X (gene ID 8239, Bethyl Laboratories, 1:1,000), rabbit anti-ERG (sc-354, polyclonal, Santa Cruz Biotechnology, 1:2,000), mouse monoclonal anti-ubiquitin (Santa Cruz Biotechnology, sc-8017, 1:1,000), and mouse anti-GAPDH (clone 258, Invitrogen, 1:2,000).

Migration assay

Costar Transwells (insert diameter 6.5 mm, 5 μm/pore) were coated with 100 μL Matrigel (Corning, 356231), covered with medium, and polymerized for 30 minutes at 30°C. The bottom chamber was filled with 600 μL of MS-5 conditioned medium. Sorted leukemia cells were allowed to migrate at 37°C for 48 hours. The transwells were removed and the cell suspension was analyzed by flow cytometer to determine the cell number and phenotype.

Gene-set enrichment analysis

Gene-set enrichment analysis (GSEA) was performed using the GSEA Java Desktop tool (v3). Gene expression levels were obtained from microarray analysis of three independent replicative experiments comparing tagBFPhigh and tagBFPneg cells. Expression levels and genes were ranked using the SAM algorithm (27, 28). The GSEA preranked tool was used to interrogate the enrichment of various expression signatures including: HSC's super enhancer (29), HSC_R (30), Ara-C resistance (31), doxorubicin resistance (32), activated β-catenin pathway (33), early T-cell development (34), targets of HOXA9 and MEIS1 up/down (35), high BAALC AML (36), epithelial-to-mesenchymal transition (EMT) up (37), and TAL1 bound and negatively/positively correlated (38). Gene-set enrichment analysis for RNA-seq data was performed using GSAASeqSP 2.0 (39).

Analysis of ERG ubiquitination

ERG ubiquitination analysis was performed essentially as described elsewhere (40) with slight modifications. Briefly, ELF153 cells were pretreated with MG132 proteasome inhibitor (5 μmol/L) and then treated with WP1130 USP9X inhibitor (5 μmol/L) for 16 hours. Cells were lysed in lysis buffer [HEPES (25 mmol/L, pH 7.5), NaCl (400 mmol/L), IGEPAL CA-630 (0.5%), DTT (1 mmol/L), glycerol (5%) and protease inhibitors]. The soluble fraction of the lysate was diluted to adjust NaCl and IGEPAL CA-630 concentrations to 100 mmol/L and 0.125%, respectively, and then precleared with magnetic Dynabeads Protein G for immunoprecipitation (Invitrogen, 10003D). Dynabeads were incubated with ERG antibody for 10 minutes at room temperature and lysates were incubated with ERG antibody-Protein G bead complexes at 4°C for 3 hours, followed by bead wash (four times) with NaCl (100 mmol/L) and IGEPAL CA-630 (0.1%). Then, immunoprecipitated complexes were boiled in Laemmli buffer for Western blot analysis. ERG ubiquitination was detected in the above prepared extracts using anti-ubiquitin antibody (Santa Cruz Biotechnology, sc-8017).

Survival analysis

Patients from different AML cohorts (41, 42) were used to determine association between USP9X levels (measured by RNA-seq RPKM analysis or microarray) and survival [overall survival (OS) and event-free survival]. Survival curves were analyzed according to the Kaplan–Meier method and compared using the log-rank test.

Chromatin immunoprecipitation sequencing gene-set enrichment analysis

Chromatin immunoprecipitation sequencing (ChIP-seq) gene-set enrichment analysis was performed using GREAT algorithm (43). ChIP-seq datasets used for the enrichment analysis were GSE49091 (44), GSE25000 (45), GSE29181 (38), and GSE50625 (46).

ChIP-seq diagrams

ChIP-seq results were extracted from the following studies: CD34+ cord blood (GSE23730; ref. 47), SKNO1 (GSE23730; ref. 47), Kasumi1 (GSE76464; ref. 48). Wig files were converted to bigwig files and processed using UCSC genome browser (49).

Protein–protein analysis

Protein–protein interaction analysis was performed using esyN platform (50).

Gene expression resources for USP9X expression analysis

USP9X expression in the different stages of hematopoietic hierarchy was extracted from GSE42414 (http://jdstemcellresearch.ca/node/32; ref. 51). USP9X expression for each AML cytogenetic group in The Cancer Genome Atlas (TCGA) was extracted from https://cancergenome.nih.gov/cancersselected/acutemyeloidleukemia. USP9X expression in chronic myeloid leukemia phases was extracted from GSE4170 (52). USP9X expression in the different AML cytogenetic-risk clusters was extracted from GSE1159 (53).

Identification and characterization of leukemia cells with heterogeneous ERG+85 reporter activities

To characterize functional heterogeneity among cells in human leukemia, we utilized stem cell enhancer element from ERG TF regulatory region (ERG+85). This element contains binding sites for numerous stem cell regulators and possesses high activity in the most primitive normal hematopoietic cells and in the subset of human bulk AML and T-ALL (20). On the basis of our findings that ERG+85 enhancer is particularly active in human HSPCs and a subset of AMLs, we recently developed a lentiviral fluorescent reporter in which ERG+85 enhancer regulates tagBFP expression, while a constitutively active EF1/SV40 promoter drives EGFP cassette (Fig. 1A; ref. 17). Using this tool, we measured ERG+85 reporter activation in a wide range of human leukemia lines and found that its activity positively correlates with the endogenous ERG protein levels (Supplementary Fig. S1; ref. 17). As ERG+85 enhancer can bind numerous high variance TFs implicated in regulation of stem cells in both T- and myeloid leukemogenesis, we hypothesized that this reporter can be a tool to track, characterize, and link transcriptional and functional heterogeneity of leukemia cells. To validate this hypothesis, we selected three leukemia lines (Jurkat, ELF-153, and Kasumi-1) in which we discovered a distinct cell population characterized by the elevated tagBFP activity as compared with the same line infected with a vector lacking ERG+85 enhancer (pMIN; Fig. 1B). Jurkat cells exhibited the lowest activation of the ERG+85 reporter, while ELF-153 and Kasumi-1 lines had the intermediate and high activities, respectively, in correlation with their respective ERG levels (Supplementary Fig. S1B).

Figure 1.

ERG+85 enhancer–based reporter demonstrates heterogeneous activity among human leukemia lines. A, Diagram of the lentiviral vector with “stemness-related TFs” (colored geometric shapes) interacting with cis elements identifiable in ERG+85 enhancer. B, Jurkat, ELF-153, and Kasumi-1 cell lines were infected with pMIN and pMIN-ERG+85 vectors. Representative EGFP and tagBFP intensity and distribution in each cell line are shown.

Figure 1.

ERG+85 enhancer–based reporter demonstrates heterogeneous activity among human leukemia lines. A, Diagram of the lentiviral vector with “stemness-related TFs” (colored geometric shapes) interacting with cis elements identifiable in ERG+85 enhancer. B, Jurkat, ELF-153, and Kasumi-1 cell lines were infected with pMIN and pMIN-ERG+85 vectors. Representative EGFP and tagBFP intensity and distribution in each cell line are shown.

Close modal

Relative chemo- and radioresistance are principal hallmarks of cancer stem cells in leukemia (54) and solid tumors (55, 56). However, the identity and the dynamics of therapy-resistant leukemia cells among the bulk malignant cells remain unclear. To test whether cells with heterogeneous ERG+85 activity differ in their response to genotoxic stress, we selected Jurkat leukemia cell line with a distinctive tagBFPHigh and tagBFPNeg subpopulations (Fig. 1B). We detected no difference in the expansion rate between tagBFPHigh and tagBFPNeg cells during in vitro culturing (Supplementary Fig. 2A). However, upon exposure of cells to ionizing radiation, we observed time-dependent enrichment in the proportion of the viable tagBFPHigh cells pointing to their relative resistance. (Fig. 2A; Supplementary Fig. S2B). FACS analysis revealed a diminished induction of the ionizing radiation (IR)-induced cell death in the tagBFPHigh cells, which is consistent with their relative resistance to this injury. (Fig. 2B; Supplementary Fig. S2C). To further elaborate these findings, we flow sorted EGFP+tagBFPHigh cells and EGFP+tagBFPNeg cells (Fig. 2C) and measured their colony-forming capacity. Analysis showed that individual tagBFPHigh cells had higher clonogenic capacity in methylcellulose as compared with the tagBFPNeg cells under both normal and irradiation conditions (Fig. 2D). Importantly, treatment of the ERG+85–expressing Jurkat cells with the additional genotoxic stressor doxorubicin led to a similar increase in the proportion of tagBFPHigh cells, suggesting an inherent resistance of this subpopulation to certain DNA-damaging agents (Fig. 2E). To validate our findings in primary samples, we infected patient-derived AML samples (n = 4) with the ERG+85 reporter followed by treatment with Ara-C or doxorubicin, both commonly used antileukemia drugs. Remarkably, we observed an elevated proportion of tagBFPHigh cells among Ara-C and doxorubicin survivors in three of four samples tested (Fig. 2F; Supplementary Fig. S2D).

Figure 2.

Functional characterization of leukemic cells expressing various levels of the ERG+85 reporter. A, Effect of ionizing radiation exposure (3 Gy) on the relative fraction of tagBFPHigh cells in the bulk Jurkat cells infected with the reporter as assessed by flow cytometry at indicated time points. n = 5 independent irradiation and recovery experiments. B, Ionizing radiation (5 Gy) induced apoptosis in tagBFPNeg and tagBFPHigh Jurkat cells at 24 hours postirradiation, quantitated using Annexin V and Zombie-NIR assay. n = 3 independent irradiation experiments. C, Gating strategy to sort tagBFPNeg and tagBFPHigh subpopulations in Jurkat cell line model. D, Clonogenic potential of tagBFPNeg and tagBFPHigh Jurkat cells exposed or not to irradiation (1 Gy; n = 3). E,ERG+85 enhancer activity analysis [measured as mean fluorescence intensity (MFI) ratio] after treatment of Jurkat cells with doxorubicin (100 nmol/L) for the indicated time (n = 3). F,ERG+85 enhancer activity analysis (measured as tagBFP mean fluorescence intensity) after treatment of AML 8227 sample with Ara-C (1 μmol/L) for 48 hours. One representative plot of three independent experiments is shown. *, P < 0.05; **, P < 0.01; ***, P < 0.001; ****, P < 0.0001.

Figure 2.

Functional characterization of leukemic cells expressing various levels of the ERG+85 reporter. A, Effect of ionizing radiation exposure (3 Gy) on the relative fraction of tagBFPHigh cells in the bulk Jurkat cells infected with the reporter as assessed by flow cytometry at indicated time points. n = 5 independent irradiation and recovery experiments. B, Ionizing radiation (5 Gy) induced apoptosis in tagBFPNeg and tagBFPHigh Jurkat cells at 24 hours postirradiation, quantitated using Annexin V and Zombie-NIR assay. n = 3 independent irradiation experiments. C, Gating strategy to sort tagBFPNeg and tagBFPHigh subpopulations in Jurkat cell line model. D, Clonogenic potential of tagBFPNeg and tagBFPHigh Jurkat cells exposed or not to irradiation (1 Gy; n = 3). E,ERG+85 enhancer activity analysis [measured as mean fluorescence intensity (MFI) ratio] after treatment of Jurkat cells with doxorubicin (100 nmol/L) for the indicated time (n = 3). F,ERG+85 enhancer activity analysis (measured as tagBFP mean fluorescence intensity) after treatment of AML 8227 sample with Ara-C (1 μmol/L) for 48 hours. One representative plot of three independent experiments is shown. *, P < 0.05; **, P < 0.01; ***, P < 0.001; ****, P < 0.0001.

Close modal

In summary, we developed a tool to identify and characterize leukemia cells with distinct stress tolerance characteristics. In addition, we defined ERG+85High cells as “stress-resistant” subset in a population of leukemia cells.

ERG+85High cells exhibit regenerative and invasive properties

To characterize the stability of cellular states identified by ERG+85 reporter, we analyzed tagBFP expression dynamics in sorted tagBFPHigh and tagBFPNeg Jurkat and ELF153 cells. FACS-assisted tagBFP analysis revealed that tagBFPHigh cells capable to regenerate the original population consisting of tagBFPHigh, tagBFPIntermediate, and tagBFPNeg cells. On the other hand, tagBFPNeg cells did not give rise to tagBFPHigh cells even over the extended period of 30 days in culture (Fig. 3A; Supplementary Fig. S3). Similar pattern of tagBFP distribution was found when we measured ERG+85 reporter activity in the progeny of an individual Jurkat tagBFPHigh or tagBFPNeg cells plated in liquid or methylcellulose colony-forming assays (Supplementary Fig. S3A and S3B).

Figure 3.

Regeneration and migration/invasion potential of tagBFPHigh and tagBFPNeg fractions. A, Flow cytometry–based analysis of tagBFP dynamics in sorted tagBFPHigh (blue histogram) and tagBFPNeg (red histogram) cells allowed to regenerate for a month. Dotted histograms represent tagBFP levels in tagBFPHigh and tagBFPNeg cells after the sort. A representative FACS plot is shown. Results are representative of eight independent experiments. B, Flow cytometric analysis of the endogenous ERG protein level in tagBFPNeg (red) and tagBFPHigh (blue) subpopulations. Pale histograms represent the isotype control staining for each fraction. A representative plot of three independent experiments is shown. C, Migration/invasion transwell assay scheme. D, Differential migration and invasion potential of tagBFPNeg and tagBFPHigh cells sorted from Jurkat and ELF153 cell lines. Mean ± SD of three (ELF153) and four (Jurkat) independent experiments is shown. **, P < 0.01; ***, P < 0.001; ns, nonsignificant.

Figure 3.

Regeneration and migration/invasion potential of tagBFPHigh and tagBFPNeg fractions. A, Flow cytometry–based analysis of tagBFP dynamics in sorted tagBFPHigh (blue histogram) and tagBFPNeg (red histogram) cells allowed to regenerate for a month. Dotted histograms represent tagBFP levels in tagBFPHigh and tagBFPNeg cells after the sort. A representative FACS plot is shown. Results are representative of eight independent experiments. B, Flow cytometric analysis of the endogenous ERG protein level in tagBFPNeg (red) and tagBFPHigh (blue) subpopulations. Pale histograms represent the isotype control staining for each fraction. A representative plot of three independent experiments is shown. C, Migration/invasion transwell assay scheme. D, Differential migration and invasion potential of tagBFPNeg and tagBFPHigh cells sorted from Jurkat and ELF153 cell lines. Mean ± SD of three (ELF153) and four (Jurkat) independent experiments is shown. **, P < 0.01; ***, P < 0.001; ns, nonsignificant.

Close modal

ERG protein–level analysis by intracellular flow cytometry and Western blotting revealed that tagBFPHigh Jurkat cells had a 2- to 3-fold higher levels of ERG protein as compared with tagBFPNeg cells (Fig. 3B; Supplementary Fig. S3C). Of note, sorted tagBFPHigh and tagBFPNeg Jurkat cells contained a similar number of viral integrations (Supplementary Fig. S3D). Collectively, these results argue against a lentivirus integration bias effect or stochastic variation in tagBFP expression. Furthermore, continuum of cells with various tagBFP levels in the progeny of the tagBFPHigh cells suggests gradual rather than on/off switch mechanism regulating ERG+85 reporter dynamics.

To determine a relative regenerative potential of tagBFPHigh and tagBFPNeg cells, we performed methycellulose colony serial replating assay. We found that tagBFPHigh ELF-153 cells gave rise to more colonies compared with tagBFPNeg cells upon the initial, secondary, and tertiary platings. Strikingly, no colonies were generated by the tagBFPNeg cells upon the third plating suggesting severely diminished self-renewal potential of this subpopulation (Supplementary Fig. S4A). To quantitate the regenerative capacity of patient-derived AML blasts expressing various levels of ERG+85 reporter, we utilized stroma-supported cultures and applied limiting dilution analysis (LDA) conditions. The LDA results revealed 1.7- to 23-fold higher frequency of culture-initiating cells in the tagBFPHigh population relative to the tagBFPNeg fraction. In agreement with this, tagBFPHigh cells were able to regenerate and demonstrated a substantial proliferative potential (up to 40-fold expansion), whereas the majority of the tagBFPNeg cell–initiated cultures failed to expand (Supplementary Fig. S4B–S4D). To better characterize the stability of these functionally distinct states, we analyzed tagBFP dynamics in sorted tagBFPHigh and tagBFPNeg primary AML cells at the day of sorting and after two weeks on stroma. Our results revealed that while tagBFPHigh cells could regenerate cells with high, intermediate, and low ERG+85 reporter activity, the tagBFPNeg blasts did not give rise to the tagBFPHigh cells (Supplementary Fig. S4B–S4D). These results obtained with primary AML samples agree with our findings in the reporter-expressing leukemia cell lines.

Elevated chemotactic migration and invasion capacities are characteristics of stem-like cells in solid tumors as well as in leukemia (57, 58). Thus, we hypothesized that leukemia cells, which are characterized by the higher levels of stem cell–specific ERG+85 enhancer, would exhibit distinct migration and invasion characteristics. To test this hypothesis, we utilized Transwell assay to examine migration of tagBFPHigh and tagBFPNeg cells toward MS5 conditioned medium and their invasion characteristics through reconstituted basement membrane (Matrigel; Fig. 3C). Sorted tagBFPHigh Jurkat cells had higher migration rate as compared with sorted tagBFPNeg cells. In the additional cell model, ELF153, we detected similar migration characteristics in sorted ELF-153 ERG+85–expressing cells. Strikingly, in Jurkat and ELF-153–based models, but not in Kasumi-1, only tagBFPHigh cells demonstrated the invasion ability through the Matrigel layer while virtually no such ability was observed in the tagBFPNeg population (Fig. 3D; Supplementary Fig. S4F). Thus, these functional assays reveal an increased invasive potential of ERG+85High cells. Furthermore, we provide evidence of ERG+85High cells’ capability to regenerate the original “phenotypic” heterogeneity. Combined with elevated stress tolerance, these functional characteristics indicate that ERG+85High cells persist preferentially in a “stem-like” state.

ERG+85-tagBFP reporter enriches for LSCs

The ability to initiate and sustain leukemia is the gold-standard assay for LSCs identification (59). To test whether ERG+85 activity and associated phenotypic heterogeneity correlate with leukemogenic potential, we performed an in vivo limiting dilution analysis (LDA). To that end, increased doses of tagBFPHigh and tagBFPNeg Jurkat cells were transplanted intrafemorally into immunodeficient recipients (Fig. 4A). Both tagBFPHigh and tagBFPNeg cells initiated local leukemic engraftment in the injected bone with the higher blast load formed by the tagBFPHigh cells (Fig. 4B).

Figure 4.

Engraftment capability of tagBFPHigh cells in NSG-mice. A, Study design for leukemia establishment and engraftment analysis. B, Leukemic engraftment in the injected bone [right femur (RF)] and in the noninjected collateral bones was analyzed 8 weeks postinjection of Jurkat cells (1 × 105 cells/mouse). Each symbol represents engraftment level in a single recipient. Horizontal lines represent mean for each group. Welch t test was used. C, Leukemia-initiating cell frequency of tagBFPHigh (black line) and tagBFPNeg (red line) Jurkat cells was determined using LDA. The dotted lines indicate the 95% confidence interval. Supplementary Table S2 contains information regarding cell dose and the number of recipients used to calculate the leukemia-initiating cell frequency. D, Flow cytometry analysis of tagBFPHigh (blue) and tagBFPNeg (red) Jurkat cells regeneration capacity in vivo. A representative FACS plot of a single mouse recipient is shown. ns, nonsignificant.

Figure 4.

Engraftment capability of tagBFPHigh cells in NSG-mice. A, Study design for leukemia establishment and engraftment analysis. B, Leukemic engraftment in the injected bone [right femur (RF)] and in the noninjected collateral bones was analyzed 8 weeks postinjection of Jurkat cells (1 × 105 cells/mouse). Each symbol represents engraftment level in a single recipient. Horizontal lines represent mean for each group. Welch t test was used. C, Leukemia-initiating cell frequency of tagBFPHigh (black line) and tagBFPNeg (red line) Jurkat cells was determined using LDA. The dotted lines indicate the 95% confidence interval. Supplementary Table S2 contains information regarding cell dose and the number of recipients used to calculate the leukemia-initiating cell frequency. D, Flow cytometry analysis of tagBFPHigh (blue) and tagBFPNeg (red) Jurkat cells regeneration capacity in vivo. A representative FACS plot of a single mouse recipient is shown. ns, nonsignificant.

Close modal

Migration and leukemia regeneration in the noninjected bones characterize LSCs with the highest initiating potential (60). Strikingly, only tagBFPHigh cells were able to repopulate collateral bones in agreement with their elevated regenerative, migratory, and invasive potentials observed in vitro. Using repopulation of the noninjected bone marrow as LDA read-out, we found an 8-fold enrichment in LSCs in tagBFPHigh cells as compared with tagBFPNeg fraction (Fig. 4C; Supplementary Table S2). Interestingly, while ERG+85 reporter activity remained negative in the leukemic engraftment formed by the tagBFPNeg cells, strong regeneration of tagBFPNeg offspring was observed in mice that received tagBFPHigh cells (Fig. 4D). This in vivo heterogeneity regeneration behavior agrees with our aforementioned in vitro findings. To assess directly whether tagBFPHigh fraction contains functional LSCs with self-renewal potential we performed serial transplantation assay using patient-derived AML sample (#120791) infected with the ERG+85 reporter. We found a gradual elevation in the proportion of the tagBFP+ cells with every round of transplantation (Supplementary Fig. S4E) that paralleled an increase in the LSC frequency (primary 1:50,000, tertiary 1:100; ref. 17). Thus, our in vivo results reveal that ERG+85 reporter–positive cells demarcate LSC-enriched cell population and LSCs properties, such as migration and invasion, are tightly associated with high ERG+85 activity.

Transcriptional profiling of ERG+85High and ERG+85Neg states

To reveal potential regulators that are responsible for the distinct functional traits of ERG+85-tagBFPHigh and ERG+85-tagBFPNeg Jurkat cells, we performed gene expression analysis on these fractions. 391 genes (366 upregulated and 25 downregulated, FDR < 0.01) were differentially expressed between ERG+85High and ERG+85Neg states, albeit with the modest fold change (2.5–4.8 fold change range; Fig. 5A; Supplementary Fig. S5A). To better understand the regulation mode of the tagBFPHigh transcriptome, we investigated enrichment for the transcription factors known to interact with ERG+85 enhancer in the regulatory regions of the differentially expressed genes. By utilizing a TF-specific ChIP-seq datasets, we revealed a strong enrichment for ERG and TAL1 factors in the upregulated genes in the ERG+85High fraction (Fig. 5B), demonstrating preferential role of these transcription factors in governing ERG+85High gene expression profile. In addition, unsupervised gene ontology analysis of regions bound by ERG in Jurkat cells highlighted pathways regulating T-cell activation and HSC function and implicated additional transcriptional regulators such as MEIS1 and β-catenin (Supplementary Fig. S5B).

Figure 5.

tagBFPHigh cells enriched with stemness and chemo-resistant gene sets. A, Statistical analysis of microarray (SAM) plot. The two parallel dashed lines are the cut-off threshold specified by the actual false discovery rate, and the total number of upregulated (red dots) and downregulated (green dots) genes are given. B, ChIP-seq enrichment analysis of the relevant TFs in tagBFPHigh signature genes. Scores are calculated using -log10(P) × (Enrichment Factor-EF)2 formula and the algorithm explained elsewhere. C, GSEA. Transcripts were ranked from top upregulated to top downregulated (in tagBFPHigh vs. tagBFPNeg fractions) and the ranked list was investigated using GSEA analysis. D, Clinical relevance of tagBFPHigh signature enrichments in different AML/T-ALL–related gene sets. AML cohort of TCGA with profiled transcriptome was used: the top 15% most expressing ERG AML patients and the 15% lowest were defined as ERGhigh and ERGlow AML patients, respectively. Database: AML, TCGA.

Figure 5.

tagBFPHigh cells enriched with stemness and chemo-resistant gene sets. A, Statistical analysis of microarray (SAM) plot. The two parallel dashed lines are the cut-off threshold specified by the actual false discovery rate, and the total number of upregulated (red dots) and downregulated (green dots) genes are given. B, ChIP-seq enrichment analysis of the relevant TFs in tagBFPHigh signature genes. Scores are calculated using -log10(P) × (Enrichment Factor-EF)2 formula and the algorithm explained elsewhere. C, GSEA. Transcripts were ranked from top upregulated to top downregulated (in tagBFPHigh vs. tagBFPNeg fractions) and the ranked list was investigated using GSEA analysis. D, Clinical relevance of tagBFPHigh signature enrichments in different AML/T-ALL–related gene sets. AML cohort of TCGA with profiled transcriptome was used: the top 15% most expressing ERG AML patients and the 15% lowest were defined as ERGhigh and ERGlow AML patients, respectively. Database: AML, TCGA.

Close modal

GSEA of the differentially expressed genes (Supplementary Table S3) revealed positive enrichment for gene signatures implicated in maintaining HSC identity (HSC super-enhancer–associated genes), HSC/LSC function (HSC_R, HOXA9 and MEIS1), chemotherapy resistance (Ara-C, doxorubicin), EMT, and early T-cell development in ERG+85High cells (Fig. 5C; Supplementary Fig. S5C). In agreement with ChIP-seq–based functional annotations, GSEA independently revealed enrichment of MEIS1, TAL1 targets, and activated β-catenin pathway in the ERG+85High cells. This analysis of ERG+85High and ERG+85Neg transcriptomes suggests that cell-to-cell variability in ERG+85 activation defines leukemia cells with distinct developmental, differentiation, and fitness potentials, which are consistent with the functional heterogeneity we described previously.

To determine the clinical relevance of the transcriptional signature identified in ERG+85High Jurkat cells, we tested its performance in the stratification of various primary AML and T-ALL datasets. We found significant enrichment of ERG+85High-specific transcripts in AML samples that expressed high levels of ERG as compared with samples expressing low levels of ERG. Strikingly, analysis of three independent studies (61–63) demonstrated that ERG+85High characteristic gene program was significantly enriched in AML and T-ALL samples obtained at relapse as compared with samples obtained at the diagnostic stage (Fig. 5D). In conclusion, our combined transcriptomics and bioinformatics analysis uncovered distinct transcriptional fingerprints associated with dynamic ERG+85High subpopulation with potential functional and clinical importance.

ERG/USP9X–positive feedback loop regulates heptad TFs network stability

Our results demonstrate that ERG+85 reporter activity varies among Jurkat cells, which is consistent with the highly dynamic nature of the TF network interacting with this genomic element (64–66). Positive and negative feedback loops that can reinforce or diminish TF expression (and thus stabilize ERG+85Neg or ERG+85High states) are general phenomenon in the developmental TF networks (67). Thus, we hypothesized that potential regulators of the heptad TFs can be differentially expressed in ERG+85High versus ERG+85Neg states. To validate this hypothesis, we performed bioinformatics analysis of potential gene-regulatory interactions between heptad TFs and differentially expressed genes and observed a number of regulatory interactions between the two groups (Fig. 6A). Furthermore, we confirmed by qRT-PCR assay differences suggested by microarrays for several genes implicated in these regulatory interactions (e.g., USP9X and SMAD3) as well as in leukemogenesis (e.g., ID2, EZH2, ALCAM, CD28) in independently sorted Jurkat subpopulations (Fig. 6B; Supplementary Fig. S5D).

Figure 6.

USP9X is an ERG target gene with a prognostic impact. A, Literature-based functional interaction analysis of the heptad TFs and genes differentially expressed between tagBFPNeg and tagBFPHigh cell subsets. esyN online tool was used to generate this interaction network. B, qRT-PCR based validation of the selected genes (SMAD3 and USP9X) predicted by microarray profiling to be differentially expressed in tagBFPNeg and tagBFPHigh subpopulations. n = 3 independent experiments. CHIP-seq datasets are listed in Materials and Methods. C,USP9X gene expression difference between ERG-expressing AMLs (the highest and the lowest 30% cutoff was used to define ERGHigh and ERGLow groups). D, Kaplan–Meier estimates of OS according to USP9X expression levels that were calculated using RNA-seq datasets from AML-TCGA study. E, Normalized expression of ERG and USP9X genes in Jurkat tagBFPNeg cells infected with the control (1074) or ERG-overexpressing (1074-ERG) vectors. Bars represent means ± SD; n = 3 qRT-PCR experiments. F,ERG+85 reporter activity analysis in tagBFPNeg Jurkat cells 3 or 10 days postinfection with control (1074, black contour plot) or ERG-overexpressing (1074-ERG, green contour plot) vectors. A representative contour plot analysis of four independent experiments is shown. G, ERG protein enrichment in USP9X regulatory region located 3-kb upstream of the transcriptional start site (TSS) was detected by querying the published ERG ChIP-seq datasets for normal CD34+ cells, SKNO1, and KASUMI-1 leukemia cell lines. ****, P < 0.0001. Welch t test was used.

Figure 6.

USP9X is an ERG target gene with a prognostic impact. A, Literature-based functional interaction analysis of the heptad TFs and genes differentially expressed between tagBFPNeg and tagBFPHigh cell subsets. esyN online tool was used to generate this interaction network. B, qRT-PCR based validation of the selected genes (SMAD3 and USP9X) predicted by microarray profiling to be differentially expressed in tagBFPNeg and tagBFPHigh subpopulations. n = 3 independent experiments. CHIP-seq datasets are listed in Materials and Methods. C,USP9X gene expression difference between ERG-expressing AMLs (the highest and the lowest 30% cutoff was used to define ERGHigh and ERGLow groups). D, Kaplan–Meier estimates of OS according to USP9X expression levels that were calculated using RNA-seq datasets from AML-TCGA study. E, Normalized expression of ERG and USP9X genes in Jurkat tagBFPNeg cells infected with the control (1074) or ERG-overexpressing (1074-ERG) vectors. Bars represent means ± SD; n = 3 qRT-PCR experiments. F,ERG+85 reporter activity analysis in tagBFPNeg Jurkat cells 3 or 10 days postinfection with control (1074, black contour plot) or ERG-overexpressing (1074-ERG, green contour plot) vectors. A representative contour plot analysis of four independent experiments is shown. G, ERG protein enrichment in USP9X regulatory region located 3-kb upstream of the transcriptional start site (TSS) was detected by querying the published ERG ChIP-seq datasets for normal CD34+ cells, SKNO1, and KASUMI-1 leukemia cell lines. ****, P < 0.0001. Welch t test was used.

Close modal

To investigate whether predicted regulatory interactions affect ERG+85-associated TF network dynamics and if so, by what mechanism, we focused on studying ERG and USP9X factors. USP9X is a deubiquitinase enzyme that can regulate stability of numerous cancer-related proteins in solid cancers, including ERG (40), MCL1 (68), and β-catenin (69). In agreement with higher expression of USP9X in ERG+85High fraction of Jurkat cells (Fig. 6B), we also revealed a higher expression of USP9X in AML samples characterized by elevated ERG levels (Fig. 6C). These findings point to a previously unrecognized positive correlation between the expressions of these two factors in human leukemia.

To provide insights into the potential involvement of USP9X in human leukemogenesis, we investigated changes in the USP9X expression in leukemia samples with established disease parameters. Using TCGA, we uncovered that high USP9X levels in patients with AML are associated with lower overall survival, higher relapse rate, and poor cytogenetics (Fig. 6D; Supplementary Fig. S6A–S6G).

To investigate the potential transcriptional regulation of USP9X gene by ERG, we ectopically expressed ERG cDNA in ERG+85Neg Jurkat cells. We observed USP9X mRNA upregulation in ERG-overexpressing cells as well as gradual tagBFP induction that becomes pronounced only after 10 days (Fig. 6E and F).

Furthermore, we revealed enrichment of ERG transcription factor in the USP9X promoter region in several AML cell lines and CD34+ cord blood cells by analyzing published ChIP- seq datasets (Fig. 6G). Interestingly, human HSCs expressed elevated USP9X levels relatively to the more differentiated subtypes (Supplementary Fig. S6H). A more comprehensive inspection revealed a canonical ETS TF DNA–binding motif in the USP9X promoter indicating a possible ERG-USP9X–positive feedback loop (Supplementary Fig. S6I).

Collectively, our results highlight the existence of the genetic link between USP9X, ERG, and ERG+85-associated stemness network and point to the functional involvement of USP9X in human leukemia.

USP9X inhibition impairs ERG+85 transcriptional network and inhibits leukemia cells’ functionality

To validate whether alteration in USP9X levels would affect ERG expression and ERG+85-tagBFP reporter activity, USP9X expression was downregulated using RNA interference approach (Fig. 7A; Supplementary Fig. S7A and S7B). USP9X knockdown by two different shRNAs (shUSP9X #2 and shUSP9X #3) led to the reduction in ERG+85 reporter activity coupled with a decrease in ERG protein levels (Fig. 7B). These results confirm that endogenous USP9X regulates ERG protein and uncovers high sensitivity of the ERG transcriptional network to changes in one of its components.

Figure 7.

Functional interrogation of the ERG/USP9X feedback loop. A, Jurkat cells were infected with the indicated viruses expressing USP9X or scramble shRNAs and then selected with puromycin (3 μg/mL) for 3 days. At day 5 after infection, cells were analyzed using qRT-PCR for USP9X mRNA reduction. n = 3 independent experiments. B,ERG+85 reporter activity and ERG protein expression analyses upon USP9X knockdown. The histograms shown are representative of three independent experiments. C, Differential sensitivity of tagBFPNeg and tagBFPHigh cells to USP9X inhibitor –WP1130. Cells were treated for 3 days with the indicated concentrations of WP1130 and analyzed for viability using FACS. Dye exclusion assay, n = 3 independent experiments. D, Susceptibility of primary AML samples to WP1130 treatment. Cells were treated for 24–72 hours with the indicated concentrations of USP9X inhibitor and then cell number and viability were determined by flow cytometry. E, Ubiquitination status of ERG after WP1130 treatment. ELF153 cells were treated with WP1130 in the presence of MG132. Immunoprecipitation was performed using anti-ERG antibody and immunoblotting was performed using anti-ubiquitin antibody. Input ERG protein is shown in the bottom image. F, Effect of WP1130 on the invasive phenotype of tagBFPHigh and tagBFPNeg subpopulations in ELF153 cells. Mean ± SD of three experiments. G, Western blot analysis of ERG protein level after treatment (16 hours) with the indicated concentrations of WP1130. H, Schematic depiction of the proposed ERG- and USP9X-positive feedback loop relationships. *, P < 0.05; **, P < 0.01; ****, P < 0.0001; ns, nonsignificant.

Figure 7.

Functional interrogation of the ERG/USP9X feedback loop. A, Jurkat cells were infected with the indicated viruses expressing USP9X or scramble shRNAs and then selected with puromycin (3 μg/mL) for 3 days. At day 5 after infection, cells were analyzed using qRT-PCR for USP9X mRNA reduction. n = 3 independent experiments. B,ERG+85 reporter activity and ERG protein expression analyses upon USP9X knockdown. The histograms shown are representative of three independent experiments. C, Differential sensitivity of tagBFPNeg and tagBFPHigh cells to USP9X inhibitor –WP1130. Cells were treated for 3 days with the indicated concentrations of WP1130 and analyzed for viability using FACS. Dye exclusion assay, n = 3 independent experiments. D, Susceptibility of primary AML samples to WP1130 treatment. Cells were treated for 24–72 hours with the indicated concentrations of USP9X inhibitor and then cell number and viability were determined by flow cytometry. E, Ubiquitination status of ERG after WP1130 treatment. ELF153 cells were treated with WP1130 in the presence of MG132. Immunoprecipitation was performed using anti-ERG antibody and immunoblotting was performed using anti-ubiquitin antibody. Input ERG protein is shown in the bottom image. F, Effect of WP1130 on the invasive phenotype of tagBFPHigh and tagBFPNeg subpopulations in ELF153 cells. Mean ± SD of three experiments. G, Western blot analysis of ERG protein level after treatment (16 hours) with the indicated concentrations of WP1130. H, Schematic depiction of the proposed ERG- and USP9X-positive feedback loop relationships. *, P < 0.05; **, P < 0.01; ****, P < 0.0001; ns, nonsignificant.

Close modal

To further substantiate the role of USP9X in regulation of leukemia cell properties, we utilized two partially selective USP9X deubiquitinase inhibitors WP1130 (70) and G9 (71) combined with leukemia lines differing in their ERG+85 reporter activity (Supplementary Fig. S7C). WP1130 as well as G9 treatment induced a dose-dependent decline in viability in all cell lines; however, leukemia lines that transactivated ERG+85 reporter (e.g., TEX, ME1 and ELF153) were more sensitive to USP9X inhibitor–mediated growth inhibition than cell lines lacking ERG+85 transactivation (e.g., K562, THP1, AML193; Fig. 7C; Supplementary Fig. S7D). To test whether MCL1 plays a major role in the preferential sensitivity of the ERG+85High cells to the USP9X inhibitor, we treated the same cell line panel with the MCL1-specific inhibitor S63845 (72). We found no correlation between the ERG+85 reporter activity and viability decline upon treatment with S63845 (Supplementary Fig. S7E), suggesting that MCL1 inhibition solely could not explain the ERG+85 activity–dependent sensitivity of cells to USP9X targeting. To extend our findings, we treated patient-derived AML samples containing ERG+85-positive fraction with WP1130 (Fig. 7D). Variable degree of growth inhibition in four of five samples was detected.

To test directly whether USP9X inhibition affects the ubiquitination status of the endogenous ERG TF in leukemia, we treated leukemia cells with WP1130 and performed ERG protein immunoprecipitation followed by immunoblotting with an antibody against ubiquitin. Indeed, treatment of ELF153 cells with WP1130 resulted in the time-dependent increase in the level of ubiquitinated endogenous ERG protein (Fig. 7E; Supplementary Fig. S8B).

We have previously shown that only ERG+85High cells exhibit strong invasive properties in vitro and in vivo. To examine whether this invasive potential can be affected by the novel ERG+85 regulator USP9X, we performed the invasion assay in the presence of WP1130. As can be seen in Fig. 7F, WP1130 exposure led to inhibition of tagBFPHigh cells’ invasion.

In agreement with our findings in the growth inhibition assays, we observed degradation of ERG protein upon WP1130 treatment in sensitive, but not resistant samples (Fig. 7G; Supplementary Fig. S8C).

To summarize, ERG regulates USP9X transcription by binding to its promoter while USP9X can stabilize ERG via deubiquitination. Interference with this feed-forward regulation via USP9X targeting can inhibit growth and invasion potential of the ERG+85-positive leukemias (Fig. 7H).

Advanced tools suitable for separating functionally distinct subpopulations of leukemia cells are obligatory for investigation and efficient targeting of LSCs. Moreover, functional characterization of LSC subpopulations can benefit from a real-time cell tracking tool to conduct diverse in vitro and in vivo studies. While in our recent study (17) we described the bioinformatics and experimental strategy to identify one such tool as a marker for stemness state (ERG+85 enhancer) and developed a fluorescent lentiviral reporter that can accurately recapitulate this enhancer endogenous activity, in this study, we advanced this reporter system for functional and molecular dissection of leukemia cell heterogeneity. Furthermore, we identified the USP9X as ERG-inducible deubiquitinase with a role in leukemia cells’ invasion and growth.

Cell-to-cell functional heterogeneity has being recognized as one of the driving forces for leukemia therapy failure and disease relapse (3, 4, 31, 73). So far, cell characterization strategies based on metabolic state labeling (74), dye efflux (75), or differential cell surface markers (73) provided solid support for the existence of the functionally distinct leukemia subpopulations in the same sample. Cell-to-cell heterogeneous activity of the ERG+85 reporter, as imaged by our system, allows dissection of the functional heterogeneity according to the stemness program–related activity. Indeed, elevated expression of normal HSC-associated transcripts in human leukemias constitutes an adverse predictor of patient survival independent of other genetic and clinical criteria (30, 76). Reduced apoptosis induction, superior chemo- and radioresistance of tagBFPHigh subpopulation, as revealed in this work, coupled with marked enrichment for the relapse-associated gene signature point to the association between the two characteristics. Moreover, adapting of this reporter labeling strategy to the high-throughput drug screen holds the potential for the early identification of small molecules capable targeting relapse-initiating cells.

Using this reporter in multiple established and primary AML samples, we showed that only ERG+85High leukemic cells are capable of giving rise to ERG+85Neg cells and sustain multiple rounds or replating. These findings suggestive of the extensive regenerative and self-renewal potential of the ERG+85High fraction agree with our previous findings with normal cord blood–derived CD34+ cells in which high ERG+85 activity characterize cells with the most immature (CD34+D38) phenotype (17). The lack of noticeable transition from ERG+85Neg cells toward ERG+85High cells in spite of prolonged in vitro or in vivo growth is surprising, giving the stochastic-state transitions previously reported in breast cancer (77). This finding may indicate that ERG+85 enhancer–based reporter enables identification of the more stable cell states that could not be identified by approaches utilizing a limited predefined set of cell surface markers or transient/metabolic labeling. Indeed, only by using an extensive repertoire of cell surface markers, distinct leukemia stem cells lineages were identified in a specific murine leukemia model driven by MLL-AF9 (78).

In addition, our findings shed light on the understanding of LSC phenotypes and permit their analysis in situ. ERG+85 reporter discovered at least two distinct self-renewing and leukemia-initiating lineages in Jurkat cells that differ in their functional capabilities. Boosted migratory/invasive activity in vitro combined with the capacity to repopulate distant bone marrow territories in vivo suggests that ERG+85High cells can have important clinical implications. For example, exploration of various bone marrow niches can endow selective protection on ERG+85High cells from chemotherapy or targeted therapy (54, 79). Preferential expression of genes related to EMT and activated β-catenin by ERG+85High cells, as discovered by our current transcriptomics analysis of Jurkat subsets, can provide potential clues on regulators implicated in invasive, migratory, and chemo-resistance phenotypes (57, 80). In addition, high expression of ALCAM (CD166) in ERG+85High cells, which is required for efficient migration and homing of HSCs, might also yield therapeutic opportunity to target these cells (81, 82).

Positive feedback loop between ERG TF and its deubiquitinase USP9X, which we revealed here, exemplifies a distinctive utilization of the differential ERG+85 reporter activity for reconstructing gene-regulatory networks implicated in leukemia cells’ functional heterogeneity. It can provide plausible molecular mechanism pertaining increased radiation/chemotherapy tolerance and activation of β-catenin pathway as we observed in ERG+85High cells. Indeed, elevated USP9X can stabilize antiapoptosis regulator MCL1 (68, 83) and β-catenin (69). USP9X's ability to reinforce ERG+85 stemness program can provide molecular support to our bioinformatics analysis that demonstrated USP9X elevation in the poor prognosis and relapsed AML samples. Importantly, we demonstrated that leukemia lines and patient AMLs are variably sensitive to WP1130 treatment. These findings advocate the usage of USP9X inhibitors (40, 84) as a promising approach in targeting leukemias that are dependent on ERG/USP9X axis.

In conclusion, our results validate a novel experimental tool to dissect leukemia cells' functional heterogeneity. We believe that insights obtained with this approach can assist in developing personalized treatments aimed to target stem-like subpopulations in human leukemia.

S. Izraeli is a consultant/advisory board member at SIGHTDX. No potential conflicts of interest were disclosed by the other authors.

Conception and design: N. Aqaqe, M. Yassin, A.A. Yassin, S. Izraeli, M. Milyavsky

Development of methodology: N. Aqaqe, M. Yassin, A.A. Yassin, S. Izraeli, M. Milyavsky

Acquisition of data (provided animals, acquired and managed patients, provided facilities, etc.): M. Yassin, A.A. Yassin, N. Ershaid, E. Kugler, E.R. Lechman, O.I. Gan, A. Mitchell, M. Milyavsky

Analysis and interpretation of data (e.g., statistical analysis, biostatistics, computational analysis): N. Aqaqe, M. Yassin, A.A. Yassin, M. Milyavsky

Writing, review, and/or revision of the manuscript: N. Aqaqe, M. Yassin, E. Kugler, S. Izraeli, M. Milyavsky

Administrative, technical, or material support (i.e., reporting or organizing data, constructing databases): N. Aqaqe, M. Yassin, N. Ershaid, C. Katz-Even, A. Zipin-Roitman, A. Mitchell, M. Milyavsky

Study supervision: E.R. Lechman, J.E. Dick, S. Izraeli, M. Milyavsky

The authors thank Dr. J.E. Pimanda (Prince of Wales Clinical School, University of New South Wales Sydney, Sydney, Australia) for providing the PGL2-ERG+85 luciferase vector, Dr. N. Donato (University of Michigan, Ann Arbor, MI) for providing G9, Dr. L. Broday (Tel Aviv University, Tel Aviv, Israel) for providing reagents and assistance with ubiquitination assay, Dr. L. Shlush (Weizmann Institute of Science, Rehovot, Israel) for providing AML samples. This work was partially supported by Israel Science Foundation (ISF 1512/14 to M. Milyavsky), Varda and Boaz Dotan Research Center in Hemato-Oncology (to M. Milyavsky and S. Izraeli), and Israel Cancer Research Fund (RCDA 14-171 to M. Milyavsky). N. Aqaqe is a recipient of a PhD scholarship from State of Israel Ministry of Science, Technology and Space. M. Yassin is a recipient of Israel Council for Higher Education PhD scholarship for minorities. This work was performed in partial fulfillment of the requirements for a PhD degree of Muhammad Yassin, Nasma Aqaqe, and Eitan Kugler, the Dr. Miriam and Sheldon G. Adelson Graduate School of Medicine, Sackler Faculty of Medicine, Tel Aviv University, Israel.

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

1.
Gilliland
DG
,
Jordan
CT
,
Felix
CA
. 
The molecular basis of leukemia
.
Hematol Am Soc Hematol Educ Progr
2004
:
80
97
.
doi: 10.1182/asheducation-2004.1.80
.
2.
Dohner
H
,
Weisdorf
DJ
,
Bloomfield
CD
. 
Acute myeloid leukemia
.
N Engl J Med
2015
;
373
:
1136
52
.
3.
Klco
JM
,
Spencer
DH
,
Miller
CA
,
Griffith
M
,
Lamprecht
TL
,
O’Laughlin
M
, et al
Functional heterogeneity of genetically defined subclones in acute myeloid leukemia
.
Cancer Cell
2014
;
25
:
379
92
.
4.
Yanagisawa
B
,
Ghiaur
G
,
Smith
BD
,
Jones
RJ
. 
Translating leukemia stem cells into the clinical setting: harmonizing the heterogeneity
.
Exp Hematol
2016
;
44
:
1130
7
.
5.
Vedi
A
,
Santoro
A
,
Dunant
CF
,
Dick
JE
,
Laurenti
E
. 
Molecular landscapes of human hematopoietic stem cells in health and leukemia
.
Ann N Y Acad Sci
2016
;
1370
:
5
14
.
6.
Greaves
M
. 
Leukaemia “firsts” in cancer research and treatment
.
Nat Rev Cancer
2016
;
16
:
163
72
.
7.
Marusyk
A
,
Almendro
V
,
Polyak
K
. 
Intra-tumour heterogeneity: a looking glass for cancer?
Nat Rev Cancer
2012
;
12
:
323
34
.
8.
Lapidot
T
,
Sirard
C
,
Vormoor
J
,
Murdoch
B
,
Hoang
T
,
Caceres-Cortes
J
, et al
A cell initiating human acute myeloid leukaemia after transplantation into SCID mice
.
Nature
1994
;
367
:
645
8
.
9.
Bonnet
D
,
Dick
JE
. 
Human acute myeloid leukemia is organized as a hierarchy that originates from a primitive hematopoietic cell
.
Nat Med
1997
;
3
:
730
7
.
10.
Pollyea
DA
,
Jordan
CT
. 
Therapeutic targeting of acute myeloid leukemia stem cells
.
Blood
2017
;
129
:
1627
35
.
11.
Gerber
JM
,
Smith
BD
,
Ngwang
B
,
Zhang
H
,
Vala
MS
,
Morsberger
L
, et al
A clinically relevant population of leukemic CD34(+)CD38(−) cells in acute myeloid leukemia
.
Blood
2012
;
119
:
3571
7
.
12.
Ran
D
,
Schubert
M
,
Taubert
I
,
Eckstein
V
,
Bellos
F
,
Jauch
A
, et al
Heterogeneity of leukemia stem cell candidates at diagnosis of acute myeloid leukemia and their clinical significance
.
Exp Hematol
2012
;
40
:
155
65
.
e1
.
13.
Guan
Y
,
Gerhard
B
,
Hogge
DEE
. 
Detection, isolation, and stimulation of quiescent primitive leukemic progenitor cells from patients with acute myeloid leukemia (AML)
.
Blood
2003
;
101
:
3142
9
.
14.
Lechman
ER
,
Gentner
B
,
Ng
SWK
,
Schoof
EM
,
van Galen
P
,
Kennedy
JA
, et al
miR-126 regulates distinct self-renewal outcomes in normal and malignant hematopoietic stem cells
.
Cancer Cell
2016
;
29
:
602
6
.
15.
Thomas
D
,
Majeti
R
. 
Biology and relevance of human acute myeloid leukemia stem cells
.
Blood
2017
;
129
:
1577
85
.
16.
Ho
TC
,
LaMere
M
,
Stevens
BM
,
Ashton
JM
,
Myers
JR
,
O’Dwyer
KM
, et al
Evolution of acute myelogenous leukemia stem cell properties after treatment and progression
.
Blood
2016
;
128
:
1671
8
.
17.
Yassin
M
,
Aqaqe
N
,
Yassin
AA
,
van Galen
P
,
Kugler
E
,
Bernstein
BE
, et al
A novel method for detecting the cellular stemness state in normal and leukemic human hematopoietic cells can predict disease outcome and drug sensitivity
.
Leukemia
2019
Jan 31.
[Epub ahead of print]. doi: 10.1038/s41375-019-0386-z
.
18.
Beck
D
,
Thoms
JA
,
Perera
D
,
Schutte
J
,
Unnikrishnan
A
,
Knezevic
K
, et al
Genome-wide analysis of transcriptional regulators in human HSPCs reveals a densely interconnected network of coding and noncoding genes
.
Blood
2013
;
122
:
e12
22
.
19.
Wilson
NK
,
Foster
SD
,
Wang
X
,
Knezevic
K
,
Schütte
J
,
Kaimakis
P
, et al
Combinatorial transcriptional control in blood stem/progenitor cells: genome-wide analysis of ten major transcriptional regulators
.
Cell Stem Cell
2010
;
7
:
532
44
.
20.
Diffner
E
,
Beck
D
,
Gudgin
E
,
Thoms
JAI
,
Knezevic
K
,
Pridans
C
, et al
Activity of a heptad of transcription factors is associated with stem cell programs and clinical outcome in acute myeloid leukemia
.
Blood
2013
;
121
:
2289
300
.
21.
Warner
JK
,
Wang
JCY
,
Takenaka
K
,
Doulatov
S
,
McKenzie
JL
,
Harrington
L
, et al
Direct evidence for cooperating genetic events in the leukemic transformation of normal human hematopoietic cells
.
Leukemia
2005
;
19
:
1794
1805
.
22
McKenzie
JL
,
Gan
OI
,
Doedens
M
,
Dick
JE
. 
Human short-term repopulating stem cells are efficiently detected following intrafemoral transplantation into NOD/SCID recipients depleted of CD122+ cells
.
Blood
2005
;
106
:
1259
61
.
23.
Hu
Y
,
Smyth
GK. ELDA
: 
Extreme limiting dilution analysis for comparing depleted and enriched populations in stem cell and other assays
.
J Immunol Methods
2009
;
347
:
70
8
.
24.
Thoms
JA
,
Birger
Y
,
Foster
S
,
Knezevic
K
,
Kirschenbaum
Y
,
Chandrakanthan
V
, et al
ERG promotes T-acute lymphoblastic leukemia and is transcriptionally regulated in leukemic cells by a stem cell enhancer
.
Blood
2011
;
117
:
7079
89
.
25.
Amendola
M
,
Venneri
MA
,
Biffi
A
,
Vigna
E
,
Naldini
L
. 
Coordinate dual-gene transgenesis by lentiviral vectors carrying synthetic bidirectional promoters
.
Nat Biotechnol
2005
;
23
:
108
16
.
26.
Milyavsky
M
,
Gan
OI
,
Trottier
M
,
Komosa
M
,
Tabach
O
,
Notta
F
, et al
A distinctive DNA damage response in human hematopoietic stem cells reveals an apoptosis-independent role for p53 in self-renewal
.
Cell Stem Cell
2010
;
7
:
186
97
.
27.
Tan
YD
,
Fornage
M
,
Fu
YX
. 
Ranking analysis of microarray data: a powerful method for identifying differentially expressed genes
.
Genomics
2006
;
88
:
846
54
.
28.
Tusher
VG
,
Tibshirani
R
,
Chu
G
. 
Significance analysis of microarrays applied to the ionizing radiation response
.
Proc Natl Acad Sci U S A
2001
;
98
:
5116
21
.
29.
van Galen
P
,
Viny
AD
,
Ram
O
,
Ryan
RJ
,
Cotton
MJ
,
Donohue
L
, et al
A multiplexed system for quantitative comparisons of chromatin landscapes
.
Molecular Cell
2016
;
61
:
170
80
.
30.
Eppert
K
,
Takenaka
K
,
Lechman
ER
,
Waldron
L
,
Nilsson
B
,
van Galen
P
, et al
Stem cell gene expression programs influence clinical outcome in human leukemia
.
Nat Med
2011
;
17
:
1086
93
.
31.
Farge
T
,
Saland
E
,
de Toni
F
,
Aroua
N
,
Hosseini
M
,
Perry
R
, et al
Chemotherapy-resistant human acute myeloid leukemia cells are not enriched for leukemic stem cells but require oxidative metabolism
.
Cancer Discov
2017
;
7
:
716
35
.
32.
Kang
HC
,
Kim
IJ
,
Park
JH
,
Shin
Y
,
Ku
JL
,
Jung
MS
, et al
Identification of genes with differential expression in acquired drug-resistant gastric cancer cells using high-density oligonucleotide microarrays
.
Clin Cancer Res
2004
;
10
:
272
84
.
33.
Kenny
PA
,
Enver
T
,
Ashworth
A
. 
Receptor and secreted targets of Wnt-1/beta-catenin signalling in mouse mammary epithelial cells
.
BMC Cancer
2005
;
5
:
3
.
doi: 10.1186/1471-2407-5-3
.
34.
Lee
MS
,
Hanspers
K
,
Barker
CS
,
Korn
AP
,
McCune
JM
. 
Gene expression profiles during human CD4+ T cell differentiation
.
Int Immunol
2004
;
16
:
1109
24
.
35.
Hess
JL
,
Bittner
CB
,
Zeisig
DT
,
Bach
C
,
Fuchs
U
,
Borkhardt
A
, et al
c-Myb is an essential downstream target for homeobox-mediated transformation of hematopoietic cells
.
Blood
2006
;
108
:
297
304
.
36.
Langer
C
,
Radmacher
MD
,
Ruppert
AS
,
Whitman
SP
,
Paschka
P
,
Mrózek
K
, et al
High BAALC expression associates with other molecular prognostic markers, poor outcome, and a distinct gene-expression signature in cytogenetically normal patients younger than 60 years with acute myeloid leukemia: a cancer and leukemia group B (CALGB) study
.
Blood
2008
;
111
:
5371
9
.
37.
Gotzmann
J
,
Fischer
AN
,
Zojer
M
,
Mikula
M
,
Proell
V
,
Huber
H
, et al
A crucial function of PDGF in TGF-β-mediated cancer progression of hepatocytes
.
Oncogene
2006
;
25
:
3170
85
.
38.
Sanda
T
,
Lawton
LN
,
Barrasa
MI
,
Fan
ZP
,
Kohlhammer
H
,
Gutierrez
A
, et al
Core transcriptional regulatory circuit controlled by the TAL1 complex in human T cell acute lymphoblastic leukemia
.
Cancer Cell
2012
;
22
:
209
21
.
39.
Xiong
Q
,
Mukherjee
S
,
Furey
TS
. 
GSAASeqSP: a toolset for gene set association analysis of RNA-seq data
.
Sci Rep
2014
;
4
:
6347
.
doi: 10.1038/srep06347
.
40.
Wang
S
,
Kollipara
RK
,
Srivastava
N
,
Li
R
,
Ravindranathan
P
,
Hernandez
E
, et al
Ablation of the oncogenic transcription factor ERG by deubiquitinase inhibition in prostate cancer
.
Proc Natl Acad Sci U S A
2014
;
111
:
4251
6
.
41.
Cancer Genome Atlas Research Network
Ley
TJ
,
Miller
C
,
Ding
L
,
Raphael
BJ
,
Mungall
AJ
, et al
Genomic and epigenomic landscapes of adult de novo acute myeloid leukemia
.
New Engl J of Med
2013
;
368
:
2059
74
.
42.
Wouters
BJ
,
Löwenberg
B
,
Erpelinck-Verschueren
CA
,
van Putten
WL
,
Valk
PJ
,
Delwel
R
. 
Double CEBPA mutations, but not single CEBPA mutations, define a subgroup of acute myeloid leukemia with a distinctive gene expression profile that is uniquely associated with a favorable outcome
.
Blood
2009
;
113
:
3088
91
.
43.
McLean
CY
,
Bristor
D
,
Hiller
M
,
Clarke
SL
,
Schaar
BT
,
Lowe
CB
, et al
GREAT improves functional interpretation of cis-regulatory regions
.
Nat Biotechnol
2010
;
28
:
495
501
.
44.
Sharma
NL
,
Massie
CE
,
Butter
F
,
Mann
M
,
Bon
H
,
Ramos-Montoya
A
, et al
The ETS family member GABPα modulates androgen receptor signalling and mediates an aggressive phenotype in prostate cancer
.
Nucleic Acids Res
2014
;
42
:
6256
69
.
45.
Palii
CG
,
Perez-Iratxeta
C
,
Yao
Z
,
Cao
Y
,
Dai
F
,
Davison
J
et al 
Differential genomic targeting of the transcription factor TAL1 in alternate haematopoietic lineages
.
EMBO J
2011
;
30
:
494
509
.
46.
Kwiatkowski
N
,
Zhang
T
,
Rahl
PB
,
Abraham
BJ
,
Reddy
J
,
Ficarro
SB
, et al
Targeting transcription regulation in cancer with a covalent CDK7 inhibitor
.
Nature
2014
;
511
:
616
20
.
47.
Martens
JH
,
Mandoli
A
,
Simmer
F
,
Wierenga
BJ
,
Saeed
S
,
Singh
AA
, et al
ERG and FLI1 binding sites demarcate targets for aberrant epigenetic regulation by AML1-ETO in acute myeloid leukemia
.
Blood
2012
;
120
:
4038
48
.
48.
Mandoli
A
,
Singh
AA
,
Prange
KHM
,
Tijchon
E
,
Oerlemans
M
,
Dirks
R
, et al
The hematopoietic transcription factors RUNX1 and ERG prevent AML1-ETO oncogene overexpression and onset of the apoptosis program in t(8;21) AMLs
.
Cell Rep
2016
;
17
:
2087
2100
.
49.
Kent
WJ
,
Sugnet
CW
,
Furey
TS
,
Roskin
KM
,
Pringle
TH
,
Zahler
AM
, et al
The human genome browser at UCSC
.
Genome Res
2002
;
12
:
996
1006
.
50.
Bean
DM
,
Heimbach
J
,
Ficorella
L
,
Micklem
G
,
Oliver
SG
,
Favrin
G
. 
esyN: Network building, sharing and publishing
.
PLoS One
2014
;
9
:
e106035
. doi:
10.1371/journal.pone.0106035
.
51.
Laurenti
E
,
Doulatov
S
,
Zandi
S
,
Plumb
I
,
Chen
J
,
April
C
, et al
The transcriptional architecture of early human hematopoiesis identifies multilevel control of lymphoid commitment
.
Nat Immunol
2013
;
14
:
756
63
.
52.
Radich
JP
,
Dai
H
,
Mao
M
,
Oehler
V
,
Schelter
J
,
Druker
B
, et al
Gene expression changes associated with progression and response in chronic myeloid leukemia
.
Proc Natl Acad Sci U S A
2006
;
103
:
2794
9
.
53.
Valk
PJ
,
Verhaak
RG
,
Beijen
MA
,
Erpelinck
CA
,
Barjesteh van Waalwijk van Doorn-Khosrovani
S
,
Boer
JM
, et al
Prognostically useful gene-expression profiles in acute myeloid leukemia
.
N Engl J Med
2004
;
350
:
1617
28
.
54.
Ishikawa
F
,
Yoshida
S
,
Saito
Y
,
Hijikata
A
,
Kitamura
H
,
Tanaka
S
, et al
Chemotherapy-resistant human AML stem cells home to and engraft within the bone-marrow endosteal region
.
Nat Biotechnol
2007
;
25
:
1315
21
.
55.
Abdullah
LN
,
Chow
EK-H
. 
Mechanisms of chemoresistance in cancer stem cells
.
Clin Transl Med
2013
;
2
:
3
.
56.
Rycaj
K
,
Tang
DG
. 
Cancer stem cells and radioresistance
.
Int J Radiat Biol
2014
;
90
:
615
21
.
57.
Stavropoulou
V
,
Kaspar
S
,
Brault
L
,
Sanders
MA
,
Juge
S
,
Morettini
S
, et al
MLL-AF9 expression in hematopoietic stem cells drives a highly invasive AML expressing EMT-related genes linked to poor outcome
.
Cancer Cell
2016
;
30
:
43
58
.
58.
Sato
R
,
Semba
T
,
Saya
H
,
Arima
Y
. 
Concise review: stem cells and epithelial-mesenchymal transition in cancer: biological implications and therapeutic targets
.
Stem Cells
2016
;
34
:
1997
2007
.
59.
Kreso
A
,
Dick
JE
. 
Evolution of the cancer stem cell model
.
Cell Stem Cell
2014
;
14
:
275
91
.
60.
Hope
KJJ
,
Jin
L
,
Dick
JEE
. 
Acute myeloid leukemia originates from a hierarchy of leukemic stem cell classes that differ in self-renewal capacity
.
Nat Immunol
2004
;
5
:
738
43
.
61.
Shlush
LI
,
Mitchell
A
,
Heisler
L
,
Abelson
S
,
Ng
SWK
,
Trotman-Grant
A
, et al
Tracing the origins of relapse in acute myeloid leukaemia to stem cells
.
Nature
2017
;
547
:
104
8
.
62.
Li
S
,
Garrett-Bakelman
FE
,
Chung
SS
,
Sanders
MA
,
Hricik
T
,
Rapaport
F
, et al
Distinct evolution and dynamics of epigenetic and genetic heterogeneity in acute myeloid leukemia
.
Nat Med
2016
;
22
:
792
9
.
63.
Clappier
E
,
Gerby
B
,
Sigaux
F
,
Delord
M
,
Touzri
F
,
Hernandez
L
, et al
Clonal selection in xenografted human T cell acute lymphoblastic leukemia recapitulates gain of malignancy at relapse
.
J Exp Med
2011
;
208
:
653
61
.
64.
Unnikrishnan
A
,
Guan
YF
,
Huang
Y
,
Beck
D
,
Thoms
JA
,
Peirs
S
, et al
A quantitative proteomics approach identifies ETV6 and IKZF1 as new regulators of an ERG-driven transcriptional network
.
Nucleic Acids Res
2016
;
44
:
10644
61
.
65.
Schutte
J
,
Wang
H
,
Antoniou
S
,
Jarratt
A
,
Wilson
NK
,
Riepsaame
J
, et al
An experimentally validated network of nine haematopoietic transcription factors reveals mechanisms of cell state stability
.
Elife
2016
;
5
:
e11469
.
66.
Wilson
NK
,
Timms
RT
,
Kinston
SJ
,
Cheng
YH
,
Oram
SH
,
Landry
JR
, et al
Gfi1 expression is controlled by five distinct regulatory regions spread over 100 kilobases, with Scl/Tal1, Gata2, PU.1, Erg, Meis1, and Runx1 acting as upstream regulators in early hematopoietic cells
.
Mol Cell Biol
2010
;
30
:
3853
63
.
67.
Wilkinson
AC
,
Nakauchi
H
,
Gottgens
B
. 
Mammalian transcription factor networks: recent advances in interrogating biological complexity
.
Cell Syst
2017
;
5
:
319
31
.
68.
Schwickart
M
,
Huang
X
,
Lill
JR
,
Liu
J
,
Ferrando
R
,
French
DM
, et al
Deubiquitinase USP9X stabilizes MCL1 and promotes tumour cell survival
.
Nature
2010
;
463
:
103
7
.
69.
Taya
S
,
Yamamoto
T
,
Kanai-Azuma
M
,
Wood
SA
,
Kaibuchi
K
. 
The deubiquitinating enzyme Fam interacts with and stabilizes beta-catenin
.
Genes Cells
1999
;
4
:
757
67
.
70.
Bartholomeusz
GA
,
Talpaz
M
,
Kapuria
V
,
Kong
LY
,
Wang
S
,
Estrov
Z
, et al
Activation of a novel Bcr/Abl destruction pathway by WP1130 induces apoptosis of chronic myelogenous leukemia cells
.
Blood
2007
;
109
:
3470
8
.
71.
Peterson
LF
,
Sun
H
,
Liu
Y
,
Potu
H
,
Kandarpa
M
,
Ermann
M
, et al
Targeting deubiquitinase activity with a novel small-molecule inhibitor as therapy for B-cell malignancies
.
Blood
2015
;
125
:
3588
97
.
72.
Kotschy
A
,
Szlavik
Z
,
Murray
J
,
Davidson
J
,
Maragno
AL
,
Le Toumelin-Braizat
G
, et al
The MCL1 inhibitor S63845 is tolerable and effective in diverse cancer models
.
Nature
2016
;
538
:
477
82
.
73.
Levine
JHH
,
Simonds
EFF
,
Bendall
SCC
,
Davis
KLL
,
Amir
ED
,
Tadmor
MDD
, et al
Data-driven phenotypic dissection of AML reveals progenitor-like cells that correlate with prognosis
.
Cell
2015
;
162
:
184
97
.
74.
Pearce
DJ
,
Taussig
D
,
Simpson
C
,
Allen
K
,
Rohatiner
AZ
,
Lister
TA
, et al
Characterization of cells with a high aldehyde dehydrogenase activity from cord blood and acute myeloid leukemia samples
.
Stem Cells
2005
;
23
:
752
60
.
75.
Moshaver
B
,
van Rhenen
A
,
Kelder
A
,
van der Pol
M
,
Terwijn
M
,
Bachas
C
, et al
Identification of a small subpopulation of candidate leukemia-initiating cells in the side population of patients with acute myeloid leukemia
.
Stem Cells
2008
;
26
:
3059
67
.
76.
Ng
SWK
,
Mitchell
A
,
Kennedy
JA
,
Chen
WC
,
McLeod
J
,
Ibrahimova
N
, et al
A 17-gene stemness score for rapid determination of risk in acute leukaemia
.
Nature
2016
;
540
:
433
7
.
77.
Friedmann-Morvinski
D
,
Verma
IM
. 
Dedifferentiation and reprogramming: origins of cancer stem cells
.
EMBO Rep
2014
;
15
:
244
53
.
78.
Guo
G
,
Luc
S
,
Marco
E
,
Lin
TW
,
Peng
C
,
Kerenyi
MA
, et al
Mapping cellular hierarchy by single-cell analysis of the cell surface repertoire
.
Cell Stem Cell
2013
;
13
:
492
505
.
79.
Gilbert
LA
,
Hemann
MT
. 
DNA damage-mediated induction of a chemoresistant niche
.
Cell
2010
;
143
:
355
66
.
80.
Yeung
J
,
Esposito
MT
,
Gandillet
A
,
Zeisig
BB
,
Griessinger
E
,
Bonnet
D
, et al
β-Catenin mediates the establishment and drug resistance of MLL leukemic stem cells
.
Cancer Cell
2010
;
18
:
606
18
.
81.
Chitteti
BR
,
Kobayashi
M
,
Cheng
Y
,
Zhang
H
,
Poteat
BA
,
Broxmeyer
HE
, et al
CD166 regulates human and murine hematopoietic stem cells and the hematopoietic niche
.
Blood
2014
;
124
:
519
29
.
82.
Jin
L
,
Hope
KJ
,
Zhai
Q
,
Smadja-Joffe
F
,
Dick
JE
. 
Targeting of CD44 eradicates human acute myeloid leukemic stem cells
.
Nat Med
2006
;
12
:
1167
74
.
83.
Trivigno
D
,
Essmann
F
,
Huber
SM
,
Rudner
J
. 
Deubiquitinase USP9x confers radioresistance through stabilization of Mcl-1
.
Neoplasia
2012
;
14
:
893
904
.
84.
Potu
H
,
Peterson
LF
,
Kandarpa
M
,
Pal
A
,
Sun
H
,
Durham
A
, et al
Usp9x regulates Ets-1 ubiquitination and stability to control NRAS expression and tumorigenicity in melanoma
.
Nat Commun
2017
;
8
:
14449
.