a novel mir-146a-pou3f2/smarca5 pathway regulates ......2020/09/24  · author manuscripts have been...

41
1 A novel miR-146a-POU3F2/SMARCA5 pathway regulates stemness and therapeutic response in glioblastoma Tiantian Cui 1 , Erica H. Bell 1 , Joseph McElroy 2 , Kevin Liu 3 , Ebin Sebastian 1 , Benjamin Johnson 1 , Pooja Manchanda Gulati 1 , Aline Paixao Becker 1 , Ashley Gray 3 , Marjolein Geurts 4 , Depika Subedi 5 , Linlin Yang 1 , Jessica Fleming 1 , Wei Meng 1 , Jill S. Barnholtz-Sloan 6 , Monica Venere 1 , Qi-En Wang 1 , Pierre A. Robe 7 , S. Jaharul Haque 1 , Arnab Chakravarti 1 1 Department of Radiation Oncology, Arthur G. James Hospital/Ohio State Comprehensive Cancer Center, Columbus, OH, 43210, USA 2 The Ohio State University Center for Biostatistics, Department of Biomedical Informatics, Columbus, OH, 43210, USA 3 The Ohio State University College of Medicine, Columbus, OH, 43210, USA 4 Erasmus Medical Center Cancer Institute, Rotterdam, the Netherlands 5 Berea College, Berea, KY, 40403, USA 6 Department of Population and Quantitative Health Sciences and Case Comprehensive Cancer Center, Case Western Reserve University School of Medicine, Cleveland, OH, USA 7 Department of Neurology and Neurosurgery, Brain Center Rudolf Magnus, University Medical Center Utrecht, Utrecht, the Netherlands Running Title: A novel miR-146a-POU3F2/SMARCA5 pathway in glioblastoma *Corresponding Author: Dr. Arnab Chakravarti Ohio State University Comprehensive Cancer Center and Richard L. Solove Research Institute Arthur G. James Cancer Hospital The Ohio State University Medical School 460 W. 10th Ave. Room D252F Columbus, Ohio 43210 Tel# 614-293-0672 Fax# 614- 293-1943 Email: [email protected] Disclosure of Potential Conflicts of Interest: No potential conflicts of interest were disclosed. on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

Upload: others

Post on 28-Jan-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

  • 1

    A novel miR-146a-POU3F2/SMARCA5 pathway regulates stemness and

    therapeutic response in glioblastoma

    Tiantian Cui1, Erica H. Bell1, Joseph McElroy2, Kevin Liu3, Ebin Sebastian1, Benjamin

    Johnson1, Pooja Manchanda Gulati1, Aline Paixao Becker1, Ashley Gray3, Marjolein

    Geurts4, Depika Subedi5, Linlin Yang1, Jessica Fleming1, Wei Meng1, Jill S.

    Barnholtz-Sloan6, Monica Venere1, Qi-En Wang1, Pierre A. Robe7, S. Jaharul Haque1,

    Arnab Chakravarti1

    1Department of Radiation Oncology, Arthur G. James Hospital/Ohio State

    Comprehensive Cancer Center, Columbus, OH, 43210, USA 2The Ohio State University Center for Biostatistics, Department of Biomedical

    Informatics, Columbus, OH, 43210, USA 3The Ohio State University College of Medicine, Columbus, OH, 43210, USA 4Erasmus Medical Center Cancer Institute, Rotterdam, the Netherlands 5Berea College, Berea, KY, 40403, USA 6Department of Population and Quantitative Health Sciences and Case

    Comprehensive Cancer Center, Case Western Reserve University School of

    Medicine, Cleveland, OH, USA 7Department of Neurology and Neurosurgery, Brain Center Rudolf Magnus,

    University Medical Center Utrecht, Utrecht, the Netherlands

    Running Title: A novel miR-146a-POU3F2/SMARCA5 pathway in glioblastoma

    *Corresponding Author:

    Dr. Arnab Chakravarti

    Ohio State University Comprehensive Cancer Center and Richard L. Solove

    Research Institute

    Arthur G. James Cancer Hospital

    The Ohio State University Medical School

    460 W. 10th Ave.

    Room D252F

    Columbus, Ohio 43210

    Tel# 614-293-0672

    Fax# 614- 293-1943

    Email: [email protected]

    Disclosure of Potential Conflicts of Interest: No potential conflicts of interest were

    disclosed.

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    mailto:[email protected]://mcr.aacrjournals.org/

  • 2

    Abstract

    Rapid tumor growth, widespread brain-invasion, and therapeutic resistance critically

    contribute to glioblastoma (GBM) recurrence and dismal patient outcomes. Although

    GBM stem cells (GSCs) are shown to play key roles in these processes, the

    molecular pathways governing the GSC phenotype (GBM-stemness) remain poorly

    defined. Here, we show that epigenetic silencing of miR-146a significantly correlated

    with worse patient outcome and importantly, miR-146a level was significantly lower in

    recurrent tumors compared with primary ones. Further, miR-146a overexpression

    significantly inhibited the proliferation and invasion of GBM patient-derived primary

    cells and increased their response to temozolomide (TMZ), both in vitro and in vivo.

    Mechanistically, miR-146a directly silenced POU3F2 and SMARCA5, two

    transcription factors that mutually regulated each other, significantly compromising

    GBM-stemness and increasing TMZ response. Collectively, our data show that miR-

    146a-POU3F2/SMARCA5 pathway plays a critical role in suppressing GBM-

    stemness and increasing TMZ-response, suggesting that POU3F2 and SMARCA5

    may serve as novel therapeutic targets in GBM.

    Implications: MicroRNA-146a predicts favorable prognosis and the miR-146a-

    POU3F2/SMARCA5 pathway is important for the suppression of stemness in GBM.

    Key words: miR-146a, POU3F2/SMARCA5, primary and recurrent GBM, therapeutic

    response, GSCs

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 3

    Introduction

    Glioblastomas (GBMs, World Health Organization grade IV gliomas) are the most

    aggressive and lethal primary malignant tumors of the central nervous system in

    adults (1). Recently the World Health Organization has classified GBM and lower

    grade gliomas based on the neomorphic mutation status of isocitrate dehydrogenase

    (IDH) 1 or 2, and irrespective of grades, gliomas with wild-type (wt) IDH1/2 have

    worse clinical outcome compared with those with mutant IDH1/2 (1). More than 90%

    of GBMs have wt-IDH1/2 and irrespective of IDH status patients with newly

    diagnosed GBM receive the standard of care with maximal safe resection, followed

    by adjuvant radiation and temozolomide (TMZ) chemotherapy (1-3). Despite the

    aggressive multi-modality treatments, the median survival remains within the range of

    12-15 months after diagnosis (4). This grim patient outcome is largely attributed to

    rapid tumor growth, invasion of vital brain structures, and evolution of intrinsic and

    acquired treatment resistance clones among other factors (4). However, we do not

    have a clear understanding of the molecular signaling networks driving disease

    progression and treatment resistance. Although a number of prognostic biomarkers,

    including promoter methylation of O6-methylguanine-DNA methyltransferase (MGMT),

    neomorphic mutation of IDH1/2, amplification and/or gain-of-function mutation of

    epidermal growth factor receptor (EGFR), loss of function mutation/deletion of TP53

    or PTEN, and CpG island methylation phenotype (CIMP) (5,6), have been identified

    and characterized, the molecular mechanisms underlying the activation of signaling

    pathways driving the treatment resistance and tumor recurrence have remained

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 4

    poorly understood posing a great challenge in developing truly effective therapeutic

    intervention. To this end, a number of studies demonstrate that a post-treatment

    increase in GSC population markedly contributes to tumor recurrence and

    therapeutic resistance mechanisms (7-9). GSCs possess potential of self-renewal

    and multipotent differentiation, and are responsible for tumor initiation and

    maintenance (7,10-13). These GSC functions are largely mediated by deregulation

    of Wnt/β-catenin (14), Notch (15), NF-κB (16), and JAK-STAT (17) signaling

    pathways among others resulting in an aberrant expression of downstream signature

    molecules that drive the resistance and recurrence mechanisms in GBM. To this end,

    we sought to identify molecular biomarkers/signatures of prognostic values and

    significance that drive treatment resistance and recurrence in GBM, and to target

    these drivers in preclinical GBM models for the development of novel therapeutic

    interventions.

    MicroRNAs (miRNAs) play regulatory roles through silencing the expression of target

    genes by post-transcriptional degradation and translational repression. Evidence

    shows that miRNAs are involved in regulating the progression and therapeutic

    resistance mechanisms in a variety of malignancies including GBM by modulating

    cancer stem cell functions (18). Previous studies have shown differential miRNA

    expression in GBM tumor tissues; moreover, tumors harboring altered miRNA

    expression had differential prognoses and treatment outcomes (19-22). Therefore,

    the differentially expressed miRNAs may serve as biomarkers and/or drivers of

    prognosis and treatment response in GBM. Very few studies have investigated the

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    https://www.sciencedirect.com/topics/medicine-and-dentistry/micrornahttp://mcr.aacrjournals.org/

  • 5

    expression of miRNA in primary and recurrent GBM (19,23). Hence, we undertook a

    high-throughput molecular profiling approach to identify miRNAs that correlate with

    clinical outcomes of patients with IDH1/2 wt GBM. We identified miR-146a as a tumor

    suppressor miRNA in GBM. Subsequently, we have identified and characterized two

    direct targets of miR-146a, POU domain, class 3, transcription factor 2 (POU3F2)

    and SWI/SNF Related, Matrix Associated, Actin Dependent Regulator of Chromatin,

    Subfamily A, Member 5 (SMARCA5), which are found to play key roles in tumor

    growth and treatment response.

    Materials and Methods

    Patient cohort, ethics statement, and RNA isolation

    We used two patient cohorts in this study. Briefly, a total of two hundred and sixty

    eight (n=268) FFPE tumor specimens without IDH1R132H mutation from patient

    cohort 1 with newly diagnosed GBM were used in the study. All of the patients

    underwent tumor biopsy or resection at the University of Utrecht (The Netherlands,

    cohort 1) from 2005 to 2014. A total of eleven pairs (n=11) of fresh frozen tumor

    specimens from patient cohort 2 with newly diagnosed GBM and matched recurrent

    GBM were used as a validation set in the study. All of the patients underwent tumor

    resection at University Hospitals of Cleveland (cohort 2) from 2008 to 2015 and

    provided written informed consent to participate in brain tumor research. The study

    was approved by The Ohio State University, University Hospital of Cleveland, and

    University Medical Center Utrecht institutional review boards.

    miRNA expression analysis and MGMT promoter methylation analyses

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 6

    Tumor samples were processed to generate miRNA expression data using the

    NanoString human v3a array, which contains 798 miRNA probes. DNA (250 ng) input

    was used per sample and the MGMT-STP27 model was used to calculate MGMT

    promoter methylation status from Illumina EPIC data (24).

    Cell lines and cell culture

    Normal human Astrocytes (NHAs) were purchased from Lonza (Walkersville, MD).

    Human GBM cell lines U251, T98G, LN229, U87MG, LN18, were purchased from

    ATCC (Manassas, VA). No further authentication was performed for NHAs and ATCC

    cell lines. U87MG/EGFRvIII cells were provided by Dr. Deliang Guo (The Ohio State

    University). Primary GBM patient-derived GBM30-luc cells were kindly provided by Dr.

    Balveen Kaur (The Ohio State University). Primary GBM patient-derived cells (08-387

    and 3359) were kindly provided by Dr. Jeremy Rich (UC San Diego) and originally

    isolated from tumor resections in accordance with approved institutional review board

    protocols. GBM12 and GBM43 were kindly provided by Mayo Clinic. These cell lines

    were authenticated by DNA profiling. U251, T98G, LN229, U87MG, LN18, and

    U87MG/EGFRvIII cells were grown in DMEM medium (Gibco) supplemented with

    10% (v/v) fetal bovine serum (FBS) (Invitrogen), 100U/ml penicillin/streptomycin

    (Sigma), and maintained in a humidified atmosphere with 5% CO2 at 37°C. GBM30,

    08-387, 3359, GBM12, and GBM43 cells were cultured in neurobasal medium with

    EGF (20ng/ml), FGF (20ng/ml), B27 (1×), and GlutaMax (1×), and Sodium Pyruvate

    (1×) in a humidified atmosphere with 5% CO2 at 37°C. All cells were periodically

    tested for mycoplasma contamination using Universal Mycoplasma Detection Kit

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 7

    (ATCC30-1012KTM).

    qRT-PCR analysis of mRNA expression and miRNA expression

    For mRNA expression analysis, total RNA was extracted from GBM cells using

    TRIzol reagent as previously described (24). The primers are listed in STable 1. For

    miRNA expression analysis, RNA (100 ng) was reverse transcribed using the

    TaqMan Advanced MicroRNA Assay kit (Applied Biosystems) with miRNA-specific

    primers. Primers of Taqman MicroRNA Assays for hsa-miR-146a (Assay ID: 000468)

    and RNU6B (Assay ID: 001093) were purchased from Invitrogen. Relative

    expression level of miR-146a was calculated by normalization with RNU6B

    expression in GBM cells. All the experiments were performed in triplicate.

    Demethylation tests, bisulfite modification, methylation-specific PCR and

    bisulfite sequencing

    GBM cells were plated and cultured on six-well plates. At 50% confluence, 10 µM

    DAC (Sigma) was added to the medium on days 1 and day 3. Cells were then

    harvested for total RNA isolation and RT-qPCR analysis. Bisulfite modification,

    methylation-specific PCR (MSP) and bisulfate sequencing (BS) were carried out as

    described previously (25). Primer pairs for MSP and BS are in STable 1.

    Western blot analysis

    Protein extraction and western blot analysis were performed as previously described

    (26). The antibodies used in this study are listed in STable 2. Polyclonal goat anti-

    rabbit antibody (Cell Signaling Technology) and Western Blotting Detection System

    (Millipore) were used for exposure.

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 8

    Transfection of miR-146a mimics and inhibitors, siRNA treatment, and Dox-

    inducible stable cells

    For functional studies, a specific miR-146 mimic/inhibitor (Invitrogen) and a

    respective negative control (Invitrogen) were transfected into the GBM cells. For

    siRNA transfection, specific POU3F2 1/2 and SMARCA5 1/2 siRNAs (50 nmol/L)

    (Dharmacon) and a negative control (Dharmacon) were used. For POU3F2

    overexpression, pcDNA3.1+/C-(K)DYK-POU3F2 plasmid (GenScript) was used. Tet-

    On shmimic-inducible miR-146a construct was purchased from Dharmacon. All

    siRNAs were transfected into cells using Lipofectamine 2000 transfection reagent

    (Invitrogen).

    Cell proliferation and invasion assays

    For cell proliferation assay, twenty-four hours post-transfection, cells (1000 cells/well)

    were seeded in 96-well plates and measured by CellTiter-Glo (Promega) after

    indicated time points as previously described (24). Cell invasion assays were

    performed as previously described (24). All the experiments were performed in

    triplicate.

    Colony formation and sphere formation assays

    GBM cells (100 cells/well) were used for colony formation assay as described

    previously (24). All the experiments were performed in triplicate. A total of 300 cells

    were mixed with neurobasal medium in a humidified atmosphere with 5% CO2 at

    37°C. The number of spheres was counted 2-3 weeks after cell seeding. All the

    experiments were performed in triplicate. To evaluate the frequency of sphere-

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 9

    forming cells (SFCf), cells were plated in 96-well plates in a limiting dilution manner

    (1, 5, 10, 20 cells/well) using FACS. The number of wells containing spheres was

    counted after 2 weeks, and the SFCf was calculated using the ELDA software

    http://bioinf.wehi.edu.au/software/elda/index.html (27).

    Cell viability assay

    Twenty-four hours post-transfection, cells (2000 cells/well) were seeded in 96-well

    plates, then treated with different dose of TMZ, 72 h after treatment, cell viability was

    measured by CellTiter-Glo (Promega).

    Apoptosis assay

    An Annexin V-fluorescence isothiocyanate (FITC) Apoptosis Detection Kit (Invitrogen)

    was used for apoptosis assay. At 72 h miR-146a overexpression, 106 cells were

    resuspended in 200 µl binding buffer with 10 µl Annexin-V-FITC and 5 µl propidium

    iodide, and incubated in the dark for another 30 min. Finally, cells were assessed

    using flow cytometric analysis. All the experiments were performed in triplicate.

    Luciferase reporter assay

    The 3’UTR of POU3F2 and SMARCA5 was synthesized, annealed, and then inserted

    into the Xho l and Not I sites of the pCheck2-reporter luciferase vector downstream of

    the stop codon of the gene for luciferase. To induce mutagenesis, the sequences

    complementary to the binding sites of miR-146a in the 3’UTR was replaced by the

    mutated sites (STable 1). The constructs were confirmed by sanger sequencing.

    Cells were co-transfected with the wild type or mutated construct, pCheck2 plasmid,

    and equal amount of negative control or miR-146a mimic. Luciferase assays were

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://bioinf.wehi.edu.au/software/elda/index.htmlhttp://mcr.aacrjournals.org/

  • 10

    performed using the Dual Luciferase Reporter Assay System (Promega).

    Xenograft tumor study

    Six to eight week old athymic nude mice were obtained from the Target Validation

    Shared Resource at the Ohio State University. All animal procedures were approved

    by the Subcommittee on Research Animal Care at The Ohio State University. To

    determine the frequency of tumor-initiating cells (TICf) using the limiting dilution

    assay, three cell doses (1 × 105, 1 × 104, 1 × 103) of each sample were injected

    subcutaneously into athymic nude mice. Mice were monitored for up to 4 weeks post-

    injection, and the tumor number per group within this period was used to calculate

    the TICf using aforementioned ELDA software(27). For the intracranial xenograft

    models, GBM cells (1×105 08-387/miR-146a/luc and GBM30/miR-146a/luc in 2 μl

    PBS), transduced with Tet-On inducible miR-146a and GFP-tagged luciferase

    plasmids, were implanted into the brain of the mouse. Doxycycline (Dox) grain-based

    diet (Thermo Fisher Scientific) was administered one day after injection. For drug

    treatment studies, mice were treated with vehicle or 20 mg/kg of TMZ resuspended in

    vehicle by oral gavage once a day for 5 days, starting at day 5 post-injection.

    Luciferin (Perkin Elmer) solution (100mg/kg) was used for imaging after tumor

    formation. The IVIS Lumina II imaging platform from the Ohio State University (OSU)

    Small Animal Imaging Core was used to detect and quantify the signal. Mice were

    sacrificed when they became moribund and the tumor tissues were harvested for

    miR-146a expression detection.

    Statistical analysis

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 11

    Analysis of Nanostring data was carried out in R (24). Cox regression was used to

    identify the association between expression of miRNAs (continuous) and overall

    survival with age as a covariable. miR-146a was then median dichotomized and the

    log rank test was employed to visualize the association between the expression and

    overall survival. Other statistical analyses were performed using the software

    package SPSS 23.0 (SPSS, Chicago, USA). Descriptive statistics, i.e., means ± SD,

    are shown on the Figures. Two sample t-tests or ANOVA were performed for data

    analysis for experiments with two groups or more than two groups’ comparisons.

    Spearman correlation analyses was applied to analyze the association between

    expression of miR-146a and POU3F2/SMARCA5. The publically available CGGA

    datasets were directly analyzed from the CGGA Data Portal at

    http://www.cgga.org.cn/. The detailed information of the RNA-Seq experiments and

    software used can be found at the CGGA Data Portal at http://www.cgga.org.cn/. P

    values were calculated two-sided. P value less than 0.05 was defined as statistically

    significant.

    Results

    Decreased miR-146a expression is associated with shorter overall survival (OS)

    of GBM patients

    To investigate a potential association between miRNA expression and GBM patient

    outcomes, miRNA expression profiling in GBM tumor specimens of a cohort of 268

    patients with wt-IDH1/2 was conducted using Nanostring v3 technology (24). miR-

    146a was identified to be one of the top miRNAs, which upon univariable analysis

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://www.cgga.org.cn/http://www.cgga.org.cn/http://mcr.aacrjournals.org/

  • 12

    (UVA) with continuous expression values (HR=0.658, 95%CI: 0.534-0.810, p

  • 13

    (data not shown), but 6 out of the remaining 8 pairs showed significantly lower

    expression of miR-146a in recurrent tumors compared with the matched primary

    tumors (Figure 1d, cohort 2, n=16, 8 pairs). Taken together, these data indicate that

    miR-146a acts as a favorable prognostic biomarker in GBMs, and miR-146a

    expression is significantly suppressed in recurrent GBM patients further suggesting

    that miR-146a may play a role in the treatment response mechanisms in GBM.

    MicroRNA-146a expression is partially regulated by promoter methylation in

    GBM

    To determine the endogenous expression of miR-146a in GBM cells, we performed

    RT-qPCR in normal human astrocytes (NHAs), 6 established GBM cell lines (U251,

    T98G, LN229, U87MG/EGFRvIII, U87MG, and LN18), and 3 patient-derived primary

    GBM cell lines (08-387, 3359, and GBM30). We found that the majority of GBM cell

    lines expressed significantly lower levels of miR-146a, except LN229, compared with

    the NHA cell line (Figure 1e). This result suggests that miR-146a is significantly

    downregulated in GBM cells, which is consistent with a previous study demonstrating

    that miR-146a expression is lower in GBM tumors compared with adjacent normal

    tissues (30). Promoter hyper-methylation leading to transcriptional silencing of

    miRNAs has been found in a variety of cancer (31). Specifically, Wang et al. have

    reported that demethylation of miR-146a promoter by 5-Aza-2’-deoxycytidine (DAC)

    correlates with delayed progression of castration resistant prostate cancer(32).

    Therefore, we explored if miR-146a was silenced by promoter methylation in GBM

    and found that DAC treatment significantly increased the miR-146a expression in 8

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 14

    out of 9 (8/9) cell lines examined, with exception of LN229 cell line that expressed a

    relatively higher endogenous level of miR-146a (Figure 1f). The methylation status of

    miR-146a was further determined by methylation specific PCR (MSP) in all 9 GBM

    cell lines. As shown in Figure 1g, 7/8 GBM cells showed partial methylation of the

    miR-146a promoter region with downregulated miR-146a expression. Consistent with

    DAC data, miR-146a promoter in LN229 cells was found to be unmethylated. To

    confirm the MSP results and further evaluate the methylation status of miR-146a in

    GBM cell lines, bisulfite sequencing (BS) was performed for 8 CpG sites (−183, −160,

    −150, −142, −138, −136, −132, and −128) of the promoter region near the

    transcription start site. Consistent with the MSP results, a high level of methylation

    was found in 7/8 cell lines with downregulated miR-146a expression (Figure 1h). The

    MSP result prompted us to analyze the methylation status of miR-146a in GBM tumor

    tissues. We selected tumor tissues with relatively low expression levels of miR-146a

    in cohort 1 and, as expected, methylation of miR-146a was found in 9/10 tumor

    tissues (Figure 1i). These results suggest that hyper-methylation of CpG islands on

    the miR-146a promoter contributes, in part, to the downregulation of miR-146a

    expression in GBM.

    Overexpression of miR-146a inhibits cell proliferation and invasion in vitro and

    in vivo

    Given that high expression of miR-146a was associated with better OS in GBM

    patients and it was downregulated in GBM cells, we hypothesized that miR-146a

    might play an important role in restraining GBM progression. To investigate the

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 15

    effects of miR-146a on cell proliferation and invasion in vitro, U87MG/EGFRvIII,

    GBM30, and 08-387 cells, which expressed low endogenous levels of miR-146a,

    were transfected with microRNA negative control (NC) or miR-146a mimic. On the

    other hand, LN229 cells, which expressed a relatively high level of miR-146a, were

    transfected with NC or a miR-146a inhibitor. We found that overexpression of miR-

    146a significantly inhibited proliferation of U87MG/EGFRvIII, GBM30, and 08-387

    cells (Figure 2a-c), which was reproduced in additional patient-derived xenografts

    (PDX) lines GBM12 and GBM43 with endogenous low expression of miR-146a

    (Figure S1a-b). In consistence with this, inhibition of miR-146a increased the

    proliferation of LN229 cells (Figure 2d). In addition to rapid proliferation, invasion is a

    defining hallmark of GBM cells (18). Accordingly, we determined the role of miR-146a

    on GBM cell invasion using a trans-well matrigel assay, which showed that

    overexpression of miR-146a suppressed invasion of GBM cells in vitro (Figure 2e,

    Figure S2a), whereas inhibition of miR-146a increased invasion (Figure 2f, Figure

    S2b). Cell proliferation and invasion data were then substantiated by using additional

    GBM cell lines (U87 and LN18, Figure S3a-c). Notably, we observed that

    overexpression of miR-146a in primary GBM cells significantly inhibited sphere

    formation, a characteristic feature of these cells (Figure 2g-2h). In consistence with

    previous studies, activation of NF-ĸB and ERK1/2 was found to be inhibited by

    overexpression of miR-146a in GBM cells (Figure S4). In order to study the role of

    miR-146a on tumor growth in vivo, we generated two tetracycline (doxycycline/Dox)

    inducible stable cell lines termed 08-387/miR-146a and GBM30/miR-146a. Our data

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 16

    show that expression of miR-146a was induced after Dox treatment and proliferation

    was inhibited in both cell lines (Figure S5a-d). Furthermore, an athymic nude

    xenograft mouse model was established, in which stable 08-387/miR-146a and

    GBM30/miR-146a cells were implanted into the brain. Decreased tumor growth was

    observed by IVIS imaging in Dox-induced group of mice that had higher expression

    of miR-146a (Figure 2i-2k) and an overall survival of these mice was prolonged as

    revealed by the Kaplan Meier plots (Figure 2l-2m). Representative H&E staining for

    the tumors are shown in Figure S5E. Dox-induced expression of miR-146a in vivo

    was further confirmed by qPCR (Figure 2n-2o). Taken together, these data suggest

    that overexpression of miR-146a inhibits tumor growth and prolongs survival of the

    tumor cell implanted mice.

    MicroRNA-146a enhances TMZ response in primary GBM models

    TMZ is the most widely used chemotherapy in GBM patients. To determine if miR-

    146a sensitizes GBM cells to TMZ, miR-146a was overexpressed in

    U87MG/EGFRvIII and 08-387 cell lines and cell viability was measured. We found

    that overexpression of miR-146a could significantly enhance TMZ-induced cell killing

    (Figure 3a-b). In addition, fewer colonies were formed in the miR-146a

    overexpressing group compared with the control group, after TMZ treatment, which

    was consistent with the outcome of the viability experiments (Figure 3c-3d). To

    further explore the effects of miR-146a on TMZ-induced apoptosis, we performed

    Annexin V apoptosis assay and measured caspase activation. As shown in Figure

    3e-3g, overexpression of miR-146a markedly enhanced TMZ-induced apoptosis in

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 17

    08-387 cells. Similar results were obtained in GBM30 cells (Figure S6a-b). These

    findings demonstrate that miR-146a overexpression enhanced TMZ response and

    potentiated TMZ-induced apoptosis in GBM cells. To further assess the function of

    miR-146a in TMZ response in vivo, an athymic nude xenograft mouse model was

    employed. Mice were randomized into 4 treatment groups, 10 mice per group: (i)

    Dox(-)/TMZ(-), (ii) Dox(-)/TMZ(+), (iii) Dox(+)/TMZ(-), and (iv) Dox(+)/TMZ(+). Mice

    were treated, as indicated, with TMZ by oral gavage for 5 days, and euthanized when

    displayed tumor-associated morbidity. Mice in Dox(-)/TMZ(-) group showed rapid

    tumor growth (23-25 days, median survival, 25 days), Dox(-)/TMZ(+) and

    Dox(+)/TMZ(-) groups initially slowed tumor growth, but tumor still grew (33-52 days

    and 33-53 days, respectively, median survival, 38 days). Mice in Dox(+)/TMZ(+)

    group had significant smaller tumors (Figure 3h) and had longer survival times (50-63

    days, median survival, 55 days, Figure 3i). Thus, miR-146a not only inhibited tumor

    growth, but also enhanced TMZ sensitivity both in vitro and in vivo in GBM.

    Overexpression of miR-146a downregulates stemness of GSCs

    GSCs are found to be primary drivers of tumor recurrence and therapeutic resistance

    (33). Because miR-146a was found to be downregulated in recurrent GBM tumors

    (Figure 1b-1d), accordingly, we hypothesized that overexpression of miR-146a would

    downregulate stemness in GSCs. To test this hypothesis, we first determined the

    expression of miR-146a in GSCs and non-GSCs (NGSCs), which were derived from

    human brain tumor specimens using CD133 selection (34). Consistent with prior

    reports, the GSC-enriched fractions expressed high level of Olig2 and Oct4 (Figure

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 18

    4a), which are markers of multipotent progenitors. qPCR data revealed that GSCs

    expressed miR-146a at markedly lower levels than the matched NGSCs, in both 08-

    387 and 3359 cells (Figure 4b). Interestingly, after a search of the GEO Profiles

    database (35), lower expression of miR-146a was also found in GBM stem-like cells

    (Gene Expression database: GSE23806, p=0.0304, Figure S7)(36), which confirmed

    our findings. Further, overexpression of miR-146a resulted in reduced proliferation of

    GSCs (Figure 4c-4d). Similarly, miR-146a overexpression reduced sphere formation

    frequency and sphere size by in vitro limiting dilution and sphere formation assays

    (Figure 4e-4h), as well as in vivo tumorigenicity by frequency of tumor-initiating cells

    (TICf) (Figureure 4i-4j), indicating that overexpression of miR-146a inhibited

    expansion of GSCs. Moreover, additional stemness markers were assessed at

    mRNA levels, which revealed that overexpression of miR-146a significantly

    downregulated expression of SALL, SOX2, c-Myc, Nestin, and Oct4 in 08-387 GSCs

    (Figure S8). These data suggest that overexpression of miR-146a reduced the

    stemness of GBM cells.

    POU3F2 and SMARCA5 are direct targets of miR-146a in GBM

    To identify putative target mRNAs of miR-146a, bioinformatics analyses were

    performed employing different miRNA target prediction tools. The neural transcription

    factor BRN2 (encoded by the POU3F2 gene), and SMARCA5 among others, were

    found as novel potential targets of miR-146a (Figure S9). Gene expression analysis

    using the TCGA datasets (http://gepia.cancer-pku.cn/) revealed that both POU3F2

    and SMARCA5 were highly expressed across different cancer types compared with

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://gepia.cancer-pku.cn/http://mcr.aacrjournals.org/

  • 19

    normal tissues (Figure S10a and Figure S11a), particularly in GBM (Figure 5a-b).

    Notably, highexpression of both of POU3F2 and SMARCA5 correlated with worse

    overall survival of glioma patients (Figure S10b and Figure S11b), and these were

    negatively correlated with miR-146a expression (Figure S10c and Figure S11c), as

    expected. To determine whether POU3F2 and SMARCA5 are putative targets,

    different GBM cells were transfected with miR-146a mimic, which resulted in

    significant reduction of both POU3F2 and SMARCA5 expression in U87MG,

    U87MG/EGFRvIII, T98G, GBM30, and 08-387 cell lines at the mRNA level (Figure

    5c-d). To confirm changes in expression of POU3F2 and SMARCA5 in the Dox-

    inducible stable 08-387/miR-146a cell model (Figure 5e), we treated 08-387/miR-

    146a cells with different dose of Dox, reduction of POU3F2 and SMARCA5 at both

    mRNA and protein levels was observed (Figure 5f-5g). To exclude the off-target effect,

    we detected expression of TRAF6, which is a known miR-146a target (37), and as

    expected, reduction of TRAF6 was observed when miR-146a was overexpressed by

    either miR-146a mimics (Figure S12a) or Dox treatment (Figure S12b). To further

    investigate whether POU3F2 and SMARCA5 are direct and specific targets of miR-

    146a, luciferase reporter containing wild-type/mutated 3’UTR of POU3F2 and

    SMARCA5 were constructed (Figure 5h). Wild-type or mutated 3’UTR reporters of

    POU3F2 and SMARCA5 were co-transfected with NC/miR-146a mimics into

    U87MG/EGFRvIII, GBM30, and 08-387 cells, respectively. Consistent reduction in

    luciferase activity by miR-146a was observed only with wild-type 3’UTR construct,

    but not with the mutant construct (Figure 5i-5j). Collectively, these data suggest that

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 20

    miR-146a suppressed POU3F2 and SMARCA5 by directly targeting their 3’-UTRs in

    GBM cells.

    Tumor suppression by miR-146a is mediated by POU3F2 and SMARCA5 and

    both proteins are positively correlated in GBM

    Once we found that tumor suppressive and TMZ-sensitizing functions of miR-146a

    were mediated by knocking down the expression of its direct targets POU3F2 and

    SMARCA5, we sought to determine the individual relative contributions of these

    stemness-related transcription factors to miR-146a functions. Among the primary

    GBM cell lines, we determined that 08-387 expressed relatively high levels of both

    POU3F2 and SMARCA5, which was consistent with a relatively low expression of

    miR-146a. We observed that 08-387 cells’ proliferation, invasion, and TMZ-response

    were significantly reduced by siRNA-mediated knockdown of either POU3F2 or

    SMARCA5 expression (Figure 6a-f). We noted that knockdown of either target was

    sufficient to produce effects on cell proliferation, invasion, and TMZ-response

    comparable to the effects produced by overexpression of miR-146a. To substantiate

    this notion, miR-146a was inhibited in LN229 cells that expressed higher level of

    miR146a (Figure 1e), which were then co-transfected with POU3F2 siRNA and/or

    SMARCA5 siRNA (Figure 6g). Subsequent functional assays demonstrated that miR-

    146a inhibition significantly induced cell proliferation and invasion, which were

    significantly reversed by knocking down either POU3F2 or SMARCA5 expression

    (Figure 6h-6i). In addition, miR-146a did not demonstrate a tumor suppressive

    phenotype when POU3F2/SMARCA5 was overexpressed (Figure S13a-c). Thus,

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 21

    either knocking down POU3F2 or SMARCA5 can mimic the effect of miR-146a in

    GBM, which provoked us to investigate the correlation between POU3F2 and

    SMARCA5 in GBM patients. Firstly, we analyzed mRNA expression of POU3F2 and

    SMARCA5 from both TCGA (https://www.cbioportal.org/) and CGGA

    (http://www.cgga.org.cn/) datasets, and found that POU3F2 and SMARCA5 were

    positively correlated (Figure S14a-b and Figure S15a-b). To confirm the finding in

    vitro, we knocked down either POU3F2 or SMARCA5 in 08-387 cells (Figure S15c-d),

    and found that expression of SMARCA5 and POU3F2 was decreased at both protein

    levels (Figure S15c-d). Collectively, these data confirmed that the tumor suppressive

    potential of miR-146a is largely mediated by the miR-146a/POU3F2/SMARCA5 axis

    (Figure 6i) in GBM.

    Discussion

    Rapid growth, invasion, and resistance to RT and TMZ are largely attributed to the

    GSC population in GBM tumors. GSCs are capable of executing these pathological

    functions through the activation of a number of signal transduction pathways,

    including the PI3Kinase, Wnt/β-catenin, Notch, NF-κB and Jak-Stat among others.

    Activation of these pathways results in aberrant expression of a variety of neural

    stem cell- and GSC-related genes, which contribute to the maintenance of

    proliferation, invasion, RT/TMZ-resistance functions in GSCs. Many of these

    regulatory genes happen to be oncogenes and tumor suppressor genes, which are

    post-transcriptionally suppressed by tumor suppressive miRNAs and onco-miRNAs,

    respectively (38). By NanoString analysis of miRNA expression in human GBM

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    https://www.cbioportal.org/http://www.cgga.org.cn/http://mcr.aacrjournals.org/

  • 22

    specimens, we have identified and validated that miR-146a expression was

    correlated with favorable patient overall survival. Its expression was partly

    suppressed by its promoter methylation in primary GBM, and further it was noted that

    expression of this miRNA was significantly lower in recurrent GBM indicating that

    miR-146a may have further significance in tumor progression or recurrence. These

    findings led to the hypothesis that miR-146 could be a tumor suppressor miRNA in

    GBM and that it might negatively control therapeutic resistance mechanisms likely by

    suppressing the GSC population and/or their functions. It is worth noting that miR-

    146a functions as a tumor suppressor miRNA in a number of cancers including

    NSCLC(39) and prostate cancer (40), and as an onco-miRNA in others that include

    cervical, head and neck, and hepatocellular cancers (41-43). Kim et al. reported that

    miR-146a was one of the miRNAs that involved in long survival GBM subclass, but

    did not provided definitive evidence (44). To test our first hypothesis, we took two

    reciprocal approaches: first, overexpression in primary PDX lines that expressed

    relatively lower levels of endogenous miR-146, and second, inhibition of miR-146a in

    cells, namely LN229 that expressed significantly higher level of miR-146. In these

    reciprocally engineered cell models, we measured cell proliferation, invasion, and

    stemness-markers in vitro, and tumor growth and overall mouse survival in

    orthotopically implanted xenograft tumors. Our results clearly suggest that miR-146a

    functions as a tumor suppressive miR in GBM. This observation is in good

    agreement with another recent report that shows that ectopic expression of miR-146a

    inhibits glioma development (45).

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 23

    Next, we tested if miR-146a sensitizes GBM to TMZ and RT. To this end, our in vitro

    data clearly indicate that miR-146 overexpression sensitized PDX lines to TMZ in

    vitro, and in vivo which, in turn, significantly prolonged the mouse survival. The PDX

    cell line models employed here, were sensitive to low dose (2Gy) of ionizing radiation

    limiting the evaluation of RT-sensitizing potential of miR-146a.

    Once we found that tumor suppressive and TMZ-sensitizing functions of miR-146a

    were mediated by knocking down the expression of its direct targets POU3F2 and

    SMARCA5, we sought to determine the relative contributions of these stemness-

    related transcription factors to miR-146a functions in GBM. POU3F2 (BRN2), a

    master regulator of neuronal differentiation, is a POU-domain transcription factor well

    described in developmental biology (46). It has been reported that inhibiting BRN2

    expression led to significantly reduced proliferation, migration, and invasion in SCLC,

    prostate cancer, as well as melanoma (47-50). Importantly, targeting BRN2 is a

    strategy to treat or prevent neuroendocrine differentiation in prostate cancer (50).

    SMARCA5 (hSNF2H) is a member of SWI/SNF family, and contains helicase and

    ATPase activities. hSNF2H promotes tumor growth in ovarian cancer and glioma

    (51,52). It has also been shown that miR-100 affects stem cell self-renewal and cell

    proliferation in part by targeting SMARCA5(53). Interestingly, Zhou et al. reported a

    reverse correlation between expression of miR-146a and SMARCA5 in a bladder

    tumor cell model (54), however, they did not provide evidence about the regulation

    between miR-146a and SMARCA5. Among the primary GBM cell lines we examined,

    08-387 expressed relatively high level of both POU3F2 and SMARCA5, which was

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 24

    consistent with a relatively low expression of miR-146a. We observed that 08-387

    cells’ proliferation, invasion, and TMZ-response were significantly reduced by siRNA-

    mediated knockdown of either POU3F2 or SMARCA5 expression. We noted, with

    surprise, that knockdown of either target was sufficient to produce effects on cell

    proliferation, invasion, and TMZ-response comparable to the effects produced by

    overexpression of miR-146a. To gain an insight of these unexpected observations,

    we conducted a series of experiments which revealed that POU3F2 and SMARCA5

    positively regulate each other in the absence of miR-146a involvement/modulation.

    The study has potential limitations. First, miR-146a expression was tested in a small

    sample size for primary and recurrent GBM, a large patient cohort is needed. Second,

    the PDX cell line models employed here, were sensitive to low dose (2Gy) of ionizing

    radiation limiting the evaluation of RT-sensitizing potential of miR-146a. Third, given

    the fact that POU3F2 is a transcriptional regulator, it might regulate expression of

    SMARCA5 through direct binding or indirectly through other targeted genes. Further

    studies are needed to investigate the mechanism of the regulation between POU3F2

    and SMARCA5 to identify the novel therapeutic strategies for GBM.

    In summary, focusing on the tumor suppressive functions of miR-146a in GBM, this

    study revealed the following novel findings: 1) Decreased miR-146a expression is

    associated with shorter OS independent of MGMT methylation status in GBM; 2)

    Promoter methylation-induced silencing of miR-146a drives tumor progression and

    therapeutic resistance in glioblastoma; 3) miR-146a inhibits the stemness of GBM

    cells via directly targeting POU3F2 and SMARCA5 whose expression were positively

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 25

    correlated in GBM. Data from our human cohorts and in vitro/vivo mechanistic

    analysis strongly implicate that miR-146a plays a significant role in the progression

    and therapeutic sensitivity in GBM, making it or POU3F2 and SMARCA5 to be

    potentially attractive and novel therapeutic targets.

    Author Contributions

    T.C. designed and performed experiments, analyzed results and wrote the

    manuscript; E.H.B. and S.J.H. designed experiments and edited the manuscript. J.M.

    analyzed Nanostring data, patient survival data and edited the manuscript. K.L., E.S.,

    A.G., M.G., D.S., L.Y. performed experiments and edited the manuscript. E.S., B.J.,

    and P.M.G performed the animal work. A.P.B. analyzed results and edited the

    manuscript. J.F., W.M., M.V. edited the manuscript. Q.E.W. analyzed results and

    edited the manuscript. P.A.R. provided patient samples and clinical annotations for

    cohort 1 and edited the manuscript. J.S.B-S. provided patient samples and clinical

    annotations for cohort 2 and edited the manuscript. A.C. coordinated the project and

    edited the manuscript.

    Acknowledgments

    We thank Dr. Jeremy Rich (UC San Diego, USA) for providing primary GBM patient-

    derived cells. We also thank The Ohio State University (OSU) Comprehensive

    Cancer Center Small Animal Imaging Core, The Ohio State University (OSU)

    Genomics Shared Resource (GSR), The Ohio State University (OSU) Target

    Validation Shared Resource (TVSR), and The Ohio State University (OSU)

    Comprehensive Cancer Center Pathology Core Facility supported in part by grant

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 26

    P30 CA016058, National Cancer Institute, Bethesda, MD. This work was also

    supported by National Cancer Institute [R01CA169368 (to A.C), R01CA11522358 (to

    A.C), R01CA1145128 (to A.C), R01CA108633 (to A.C), R01CA188228 (to A.C., R.B.,

    K.L., and J.S.B-S.), 1RC2CA148190 (to A.C), and U10CA180850-01 (to A.C.)]; A

    Brain Tumor Funders Collaborative Grant (to A.C.); Ohio State University

    Comprehensive Cancer Center Award (to A.C.), and the T&P Bohnenn Fund for

    Neuro-Oncology Research (grant to P.A.R.).

    References

    1. Louis DN, Perry A, Reifenberger G, von Deimling A, Figarella-Branger D,

    Cavenee WK, et al. The 2016 World Health Organization Classification of

    Tumors of the Central Nervous System: a summary. Acta Neuropathol

    2016;131(6):803-20 doi 10.1007/s00401-016-1545-1.

    2. Stupp R, Hegi ME, Mason WP, van den Bent MJ, Taphoorn MJ, Janzer RC, et

    al. Effects of radiotherapy with concomitant and adjuvant temozolomide

    versus radiotherapy alone on survival in glioblastoma in a randomised phase

    III study: 5-year analysis of the EORTC-NCIC trial. Lancet Oncol

    2009;10(5):459-66 doi 10.1016/S1470-2045(09)70025-7.

    3. Ryken TC, Kalkanis SN, Buatti JM, Olson JJ. The role of cytoreductive

    surgery in the management of progressive glioblastoma. J Neuro-Oncol

    2014;118(3):479-88 doi 10.1007/s11060-013-1336-7.

    4. Ostrom QT, Cioffi G, Gittleman H, Patil N, Waite K, Kruchko C, et al. CBTRUS

    Statistical Report: Primary Brain and Other Central Nervous System Tumors

    Diagnosed in the United States in 2012-2016. Neuro Oncol

    2019;21(Supplement_5):v1-v100 doi 10.1093/neuonc/noz150.

    5. Weller M, Pfister SM, Wick W, Hegi ME, Reifenberger G, Stupp R. Molecular

    neuro-oncology in clinical practice: a new horizon. Lancet Oncol

    2013;14(9):e370-9 doi 10.1016/S1470-2045(13)70168-2.

    6. Thakkar JP, Dolecek TA, Horbinski C, Ostrom QT, Lightner DD, Barnholtz-

    Sloan JS, et al. Epidemiologic and molecular prognostic review of

    glioblastoma. Cancer Epidemiol Biomarkers Prev 2014;23(10):1985-96 doi

    10.1158/1055-9965.EPI-14-0275.

    7. Chen J, Li YJ, Yu TS, McKay RM, Burns DK, Kernie SG, et al. A restricted cell

    population propagates glioblastoma growth after chemotherapy. Nature

    2012;488(7412):522-+ doi 10.1038/nature11287.

    8. Auffinger B, Tobias AL, Han Y, Lee G, Guo D, Dey M, et al. Conversion of

    differentiated cancer cells into cancer stem-like cells in a glioblastoma model

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 27

    after primary chemotherapy. Cell Death Differ 2014;21(7):1119-31 doi

    10.1038/cdd.2014.31.

    9. Tamura K, Aoyagi M, Wakimoto H, Ando N, Nariai T, Yamamoto M, et al.

    Accumulation of CD133-positive glioma cells after high-dose irradiation by

    Gamma Knife surgery plus external beam radiation. J Neurosurg

    2010;113(2):310-8 doi 10.3171/2010.2.JNS091607.

    10. De Bacco F, D'Ambrosio A, Casanova E, Orzan F, Neggia R, Albano R, et al.

    MET inhibition overcomes radiation resistance of glioblastoma stem-like cells.

    EMBO Mol Med 2016;8(5):550-68 doi 10.15252/emmm.201505890.

    11. Lathia JD, Mack SC, Mulkearns-Hubert EE, Valentim CLL, Rich JN. Cancer

    stem cells in glioblastoma. Gene Dev 2015;29(12):1203-17 doi

    10.1101/gad.261982.115.

    12. Singh SK, Hawkins C, Clarke ID, Squire JA, Bayani J, Hide T, et al.

    Identification of human brain tumour initiating cells. Nature

    2004;432(7015):396-401 doi 10.1038/nature03128.

    13. Liu G, Yuan X, Zeng Z, Tunici P, Ng H, Abdulkadir IR, et al. Analysis of gene

    expression and chemoresistance of CD133+ cancer stem cells in

    glioblastoma. Mol Cancer 2006;5:67 doi 10.1186/1476-4598-5-67.

    14. Zhang N, Wei P, Gong A, Chiu WT, Lee HT, Colman H, et al. FoxM1 promotes

    beta-catenin nuclear localization and controls Wnt target-gene expression

    and glioma tumorigenesis. Cancer Cell 2011;20(4):427-42 doi

    10.1016/j.ccr.2011.08.016.

    15. Wang JL, Wakeman TP, Lathia JD, Hjelmeland AB, Wang XF, White RR, et al.

    Notch Promotes Radioresistance of Glioma Stem Cells. Stem Cells

    2010;28(1):17-28 doi 10.1002/stem.261.

    16. Kim SH, Ezhilarasan R, Phillips E, Gallego-Perez D, Sparks A, Taylor D, et al.

    Serine/Threonine Kinase MLK4 Determines Mesenchymal Identity in Glioma

    Stem Cells in an NF-kappaB-dependent Manner. Cancer Cell 2016;29(2):201-

    13 doi 10.1016/j.ccell.2016.01.005.

    17. Cao Y, Lathia JD, Eyler CE, Wu Q, Li Z, Wang H, et al. Erythropoietin

    Receptor Signaling Through STAT3 Is Required For Glioma Stem Cell

    Maintenance. Genes Cancer 2010;1(1):50-61 doi

    10.1177/1947601909356352.

    18. Beyer S, Fleming J, Meng W, Singh R, Haque SJ, Chakravarti A. The Role of

    miRNAs in Angiogenesis, Invasion and Metabolism and Their Therapeutic

    Implications in Gliomas. Cancers (Basel) 2017;9(7) doi

    10.3390/cancers9070085.

    19. Matos B, Bostjancic E, Matjasic A, Popovic M, Glavac D. Dynamic expression

    of 11 miRNAs in 83 consecutive primary and corresponding recurrent

    glioblastoma: correlation to treatment, time to recurrence, overall survival and

    MGMT methylation status. Radiol Oncol 2018;52(4):422-32 doi 10.2478/raon-

    2018-0043.

    20. Huang SW, Ali ND, Zhong L, Shi J. MicroRNAs as biomarkers for human

    glioblastoma: progress and potential. Acta Pharmacol Sin 2018;39(9):1405-13

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 28

    doi 10.1038/aps.2017.173.

    21. Henriksen M, Johnsen KB, Andersen HH, Pilgaard L, Duroux M. MicroRNA

    Expression Signatures Determine Prognosis and Survival in Glioblastoma

    Multiforme-a Systematic Overview. Mol Neurobiol 2014;50(3):896-913 doi

    10.1007/s12035-014-8668-y.

    22. Low SYY, Ho YK, Too HP, Yap CT, Ng WH. MicroRNA as potential modulators

    in chemoresistant high-grade gliomas. J Clin Neurosci 2014;21(3):395-400

    doi 10.1016/j.jocn.2013.07.033.

    23. Bo LJ, Wei B, Li ZH, Wang ZF, Gao Z, Miao Z. Bioinformatics analysis of

    miRNA expression profile between primary and recurrent glioblastoma. Eur

    Rev Med Pharmaco 2015;19(19):3579-86.

    24. Cui T, Bell EH, McElroy J, Becker AP, Gulati PM, Geurts M, et al. miR-4516

    predicts poor prognosis and functions as a novel oncogene via targeting

    PTPN14 in human glioblastoma. Oncogene 2019;38(16):2923-36 doi

    10.1038/s41388-018-0601-9.

    25. Cui T, Chen Y, Yang L, Knosel T, Zoller K, Huber O, et al. DSC3 expression is

    regulated by p53, and methylation of DSC3 DNA is a prognostic marker in

    human colorectal cancer. Brit J Cancer 2011;104(6):1013-9 doi

    10.1038/bjc.2011.28.

    26. Cui TT, Chen Y, Yang LL, Knosel T, Huber O, Pacyna-Gengelbach M, et al.

    The p53 target gene desmocollin 3 acts as a novel tumor suppressor through

    inhibiting EGFR/ERK pathway in human lung cancer. Carcinogenesis

    2012;33(12):2326-33 doi 10.1093/carcin/bgs273.

    27. Hu YF, Smyth GK. ELDA: Extreme limiting dilution analysis for comparing

    depleted and enriched populations in stem cell and other assays. J Immunol

    Methods 2009;347(1-2):70-8 doi 10.1016/j.jim.2009.06.008.

    28. Stupp R, Mason WP, van den Bent MJ, Weller M, Fisher B, Taphoorn MJ, et al.

    Radiotherapy plus concomitant and adjuvant temozolomide for glioblastoma.

    N Engl J Med 2005;352(10):987-96 doi 10.1056/NEJMoa043330.

    29. Hegi ME, Diserens AC, Gorlia T, Hamou MF, de Tribolet N, Weller M, et al.

    MGMT gene silencing and benefit from temozolomide in glioblastoma. N Engl

    J Med 2005;352(10):997-1003 doi 10.1056/NEJMoa043331.

    30. Hu HQ, Sun LG, Guo WJ. Decreased miRNA-146a in glioblastoma multiforme

    and regulation of cell proliferation and apoptosis by target Notch1. Int J Biol

    Marker 2016;31(3):E270-E5 doi 10.5301/jbm.5000194.

    31. Ha M, Kim VN. Regulation of microRNA biogenesis. Nat Rev Mol Cell Biol

    2014;15(8):509-24 doi 10.1038/nrm3838.

    32. Wang XL, Gao HT, Ren LX, Gu JF, Zhang YP, Zhang Y. Demethylation of the

    miR-146a promoter by 5-Aza-2 '-deoxycytidine correlates with delayed

    progression of castration-resistant prostate cancer. Bmc Cancer 2014;14 doi

    Artn 30810.1186/1471-2407-14-308.

    33. Osuka S, Van Meir EG. Overcoming therapeutic resistance in glioblastoma:

    the way forward. J Clin Invest 2017;127(2):415-26 doi 10.1172/Jci89587.

    34. Dermawan JKT, Hitomi M, Silver DJ, Wu QL, Sandlesh P, Sloan AE, et al.

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 29

    Pharmacological Targeting of the Histone Chaperone Complex FACT

    Preferentially Eliminates Glioblastoma Stem Cells and Prolongs Survival in

    Preclinical Models. Cancer Res 2016;76(8):2432-42 doi 10.1158/0008-

    5472.Can-15-2162.

    35. Barrett T, Troup DB, Wilhite SE, Ledoux P, Rudnev D, Evangelista C, et al.

    NCBI GEO: mining tens of millions of expression profiles--database and tools

    update. Nucleic Acids Res 2007;35(Database issue):D760-5 doi

    10.1093/nar/gkl887.

    36. Schulte A, Gunther HS, Phillips HS, Kemming D, Martens T, Kharbanda S, et

    al. A distinct subset of glioma cell lines with stem cell-like properties reflects

    the transcriptional phenotype of glioblastomas and overexpresses CXCR4 as

    therapeutic target. Glia 2011;59(4):590-602 doi 10.1002/glia.21127.

    37. Stickel N, Prinz G, Pfeifer D, Hasselblatt P, Schmitt-Graeff A, Follo M, et al.

    MiR-146a regulates the TRAF6/TNF-axis in donor T cells during GVHD.

    Blood 2014;124(16):2586-95 doi 10.1182/blood-2014-04-569046.

    38. Esquela-Kerscher A, Slack FJ. Oncomirs - microRNAs with a role in cancer.

    Nat Rev Cancer 2006;6(4):259-69 doi 10.1038/nrc1840.

    39. Li YL, Wang J, Zhang CY, Shen YQ, Wang HM, Ding L, et al. MiR-146a-5p

    inhibits cell proliferation and cell cycle progression in NSCLC cell lines by

    targeting CCND1 and CCND2. Oncotarget 2016;7(37):59287-98 doi

    10.18632/oncotarget.11040.

    40. Huang Y, Tao T, Liu C, Guan H, Zhang G, Ling Z, et al. Upregulation of miR-

    146a by YY1 depletion correlates with delayed progression of prostate cancer.

    Int J Oncol 2017;50(2):421-31 doi 10.3892/ijo.2017.3840.

    41. Wang X, Tang S, Le SY, Lu R, Rader JS, Meyers C, et al. Aberrant expression

    of oncogenic and tumor-suppressive microRNAs in cervical cancer is required

    for cancer cell growth. PLoS One 2008;3(7):e2557 doi

    10.1371/journal.pone.0002557.

    42. Hung PS, Chang KW, Kao SY, Chu TH, Liu CJ, Lin SC. Association between

    the rs2910164 polymorphism in pre-mir-146a and oral carcinoma progression.

    Oral Oncol 2012;48(5):404-8 doi 10.1016/j.oraloncology.2011.11.019.

    43. Xu T, Zhu Y, Wei QK, Yuan Y, Zhou F, Ge YY, et al. A functional polymorphism

    in the miR-146a gene is associated with the risk for hepatocellular carcinoma.

    Carcinogenesis 2008;29(11):2126-31 doi 10.1093/carcin/bgn195.

    44. Kim TM, Huang W, Park R, Park PJ, Johnson MD. A Developmental

    Taxonomy of Glioblastoma Defined and Maintained by MicroRNAs. Cancer

    Res 2011;71(9):3387-99 doi 10.1158/0008-5472.Can-10-4117.

    45. Mei J, Bachoo R, Zhang CL. MicroRNA-146a inhibits glioma development by

    targeting Notch1. Mol Cell Biol 2011;31(17):3584-92 doi 10.1128/MCB.05821-

    11.

    46. Dominguez MH, Ayoub AE, Rakic P. POU-III transcription factors (Brn1, Brn2,

    and Oct6) influence neurogenesis, molecular identity, and migratory

    destination of upper-layer cells of the cerebral cortex. Cereb Cortex

    2013;23(11):2632-43 doi 10.1093/cercor/bhs252.

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 30

    47. Ishii J, Sato H, Sakaeda M, Shishido-Hara Y, Hiramatsu C, Kamma H, et al.

    POU domain transcription factor BRN2 is crucial for expression of ASCL1,

    ND1 and neuroendocrine marker molecules and cell growth in small cell lung

    cancer. Pathol Int 2013;63(3):158-68 doi 10.1111/pin.12042.

    48. Boyle GM, Woods SL, Bonazzi VF, Stark MS, Hacker E, Aoude LG, et al.

    Melanoma cell invasiveness is regulated by miR-211 suppression of the

    BRN2 transcription factor. Pigm Cell Melanoma R 2011;24(3):525-37 doi

    10.1111/j.1755-148X.2011.00849.x.

    49. Pinner S, Jordan P, Sharrock K, Bazley L, Collinson L, Marais R, et al.

    Intravital Imaging Reveals Transient Changes in Pigment Production and

    Brn2 Expression during Metastatic Melanoma Dissemination. Cancer Res

    2009;69(20):7969-77 doi 10.1158/0008-5472.Can-09-0781.

    50. Bishop JL, Thaper D, Vahid S, Davies A, Ketola K, Kuruma H, et al. The

    Master Neural Transcription Factor BRN2 Is an Androgen Receptor-

    Suppressed Driver of Neuroendocrine Differentiation in Prostate Cancer.

    Cancer Discov 2017;7(1):54-71 doi 10.1158/2159-8290.Cd-15-1263.

    51. Sheu JJ, Choi JH, Yildiz I, Tsai FJ, Shaul Y, Wang TL, et al. The roles of

    human sucrose nonfermenting protein 2 homologue in the tumor-promoting

    functions of Rsf-1. Cancer Res 2008;68(11):4050-7 doi 10.1158/0008-

    5472.CAN-07-3240.

    52. Zhao XC, An P, Wu XY, Zhang LM, Long B, Tian Y, et al. Overexpression of

    hSNF2H in glioma promotes cell proliferation, invasion, and chemoresistance

    through its interaction with Rsf-1. Tumor Biol 2016;37(6):7203-12 doi

    10.1007/s13277-015-4579-4.

    53. Deng L, Shang L, Bai S, Chen J, He X, Martin-Trevino R, et al. MicroRNA100

    inhibits self-renewal of breast cancer stem-like cells and breast tumor

    development. Cancer Res 2014;74(22):6648-60 doi 10.1158/0008-5472.CAN-

    13-3710.

    54. Zhou W, Thiery JP. Loss of Git2 induces epithelial-mesenchymal transition by

    miR146a-Cnot6L-controlled expression of Zeb1. J Cell Sci 2013;126(Pt

    12):2740-6 doi 10.1242/jcs.126367.

    Figure Legends

    Figure 1 Expression of miR-146a in GBM patient samples and cell lines. (a)

    Correlation between miR-146a expression and OS by Kaplan-Meier analysis of GBM

    patients with high (n=134) and low (n=134) expression of miR-146a. (b) Expression

    of miR-146a in primary and recurrent GBM by Nanostring analysis. (c) Expression of

    miR-146a in primary and recurrent GBM by qRT-PCR in cohort 1(14 pairs). n=3,

    **P

  • 31

    RT-qPCR in cohort 2 (8 pairs). n=3, *P

  • 32

    or GBM30/miR-146a/luc cells without or with Dox (n = 10/group). (n-o) Expression of

    miR-146a was detected after isolating RNAs from tumors harvested from mice brain

    when the mice became moribund. n=5, ***P

  • 33

    vitro limiting dilution, cells were plated in a limiting dilution manner in 96-well plates,

    and number of wells containing spheres calculated after 2-3 weeks to calculate the

    SFCf (e,g). Sphere formation assay was conducted and number of spheres was

    counted in 3 different areas and mean value was calculated (f,h). n=3, **P

  • 34

    GBM30, and 08-387 cells were co-transfected with the luciferase reporter

    constructs containing 3′UTR (wt) or 3′UTR (mut) together with NC or miR-146a

    mimics. Seventy-two hours post transfection, luciferase activity was measured and

    normalized. n = 3, *P

  • Primary Recurrent

    7.5

    7.0

    6.5

    6.0

    5.5

    5.0 p=0.0047 No

    rmali

    zed

    miR

    -146

    a L

    ev

    el

    (Nan

    os

    trin

    g)

    a b

    e f

    U251

    M U

    T98G

    M U

    U87MG/EGFRvIII

    M U

    U87MG

    M U

    LN18

    M U

    08-387

    M U

    3359

    M U

    GBM30

    M U

    LN229

    M U

    g

    Case 1

    M U

    Case 2

    M U

    Case 3

    M U

    Case 4

    M U

    Case 5

    M U

    Case 6

    M U

    Case 7

    M U M U M U M U

    Case 8 Case 9 Case 10

    h

    i

    c

    Cohort 1

    **

    ***

    ***

    **

    *** ***

    *** ***

    ***

    ** **

    d

    Cohort 2

    ** ** ** **

    * *

    *** ***

    **

    ***

    ***

    ***

    ***

    **

    ** **

    CpG sites -183 -160 -150 -142 -138 -136 -132 -128

    U87MG

    GBM30

    LN18

    U87MG/EGFRvIII

    U251

    08-387

    T98G

    3359

    LN229

    Promoter Region

    Figure 1

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • GBM30

    08-387

    NC miR-146a

    GBM30/miR-146a/luc

    a b c d

    e f g h

    i

    j k

    Dox (+) Dox (-)

    10

    17

    25

    Dox (+) Dox (-)

    Radiance

    (p/s/cm2/sr)

    2.0

    1.5

    1.0

    0.5

    0

    1×107

    08-387/miR-146a/luc GBM30/miR-146a/luc

    m

    08-387/miR-146a/luc

    LN229

    ***

    ***

    *** ***

    *** *** ***

    **

    ** **

    **

    * **

    ***

    ** p

  • miR-NC miR-146a

    DMSO

    TMZ

    (100µM)

    Annexin V

    PI

    DMSO TMZ 100µM

    NC

    miR-146a

    Dox (-) TMZ (-)

    Dox (-) TMZ(+)

    Dox (+) TMZ(-)

    Dox (+) TMZ(+)

    3.0

    2.0

    1.0

    1×106

    4.0

    5.0

    6.0

    Radiance

    (p/s/cm2/sr)

    a b c d

    e f

    h

    **

    ***

    **g

    c-caspase 3

    08-387

    TMZ(µM)

    c-parp

    parp

    caspase 3

    U87MG/EGFRvIII

    NC miR-146a

    0 100 500 0 100 500

    NC miR-146a

    0 50 100 0 50 100

    i

    08-387/miR-146a/luc

    ** p

  • Cell number

    No. of Tumor

    (injection sites)

    NC miR-146a

    2× 105 4(4) 4(4)

    2× 104 4(4) 3(4)

    2× 103 3(4) 2(4)

    TICf 1/1443 1/9158

    TICf

    (TIC/104) 6.93 1.09

    p 0.0296 08-387GSCs

    e

    i j

    g

    TICf

    Fra

    cti

    on

    of

    wells w

    ith

    ou

    t

    tu

    mo

    r s

    ph

    ere

    s (

    log

    2)

    Number of Cells

    miR-146a

    08-387GSCs

    NC

    Fra

    cti

    on

    of

    wells w

    ith

    ou

    t

    tu

    mo

    r s

    ph

    ere

    s (

    log

    2)

    Number of Cells

    miR-146a

    3359GSCs

    NC

    Group NC miR-146a

    Estimate

    (Range)

    1 in 12.7

    (9.44-17.0)

    1 in 21.8

    (15.51-30.7)

    p 0.0145

    Group NC miR-146a

    Estimate

    (Range)

    1 in 15.6

    (11.4-21.3)

    1 in 31.4

    (21.3-46.4)

    p 0.00457

    SFCf

    SFCf

    *

    f

    h

    **

    **

    OCT4

    OLIG2

    08-387 3359

    GAPDH

    a b c d

    *

    **

    ***

    ***

    (Days)

    Figure 4

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • Dox 1.0µM - +

    SMARCA5

    08-387/miR146a

    GAPDH

    POU3F2

    GBM

    6

    5

    4

    3

    2

    1

    0

    PO

    U3F

    2 E

    xp

    res

    sio

    n

    (TC

    GA

    , lo

    g)

    T N GBM

    6

    5

    4

    3

    2

    1

    0

    SM

    AR

    CA

    5 E

    xp

    res

    sio

    n

    (TC

    GA

    , lo

    g)

    T N

    a b

    e

    c

    d

    f g h

    i j

    ** ** **

    *

    **

    ***

    ***

    **

    **

    ** **

    * * * *

    **

    *

    **

    ***

    **

    ** *

    *

    ** **

    *

    Figure 5

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • 0 50 100 0 50 100

    siCtrl siSMARCA5

    TMZ(µM)

    GAPDH

    parp

    c-parp

    SMARCA5

    0 50 100 0 50 100

    siCtrl siPOU3F2

    TMZ(µM)

    GAPDH

    parp

    c-parp

    POU3F2

    a

    c d e

    POU3F2

    GAPDH

    SMARCA5

    GAPDH

    POU3F2

    SMARCA5

    GAPDH

    miR-146a inhibitor

    siPOU3F2

    siSMARCA5

    + - + + +

    - - + - +

    - - - + +

    f g

    h

    ** ** **

    **

    miR-146a inhibitor

    siPOU3F2

    siSMARCA5

    + - + + +

    - - + - + - - - + +

    + - + + +

    - - + - + - - - + +

    **

    **

    + - + + +

    - - + - + - - - + +

    miR-146a inhibitor

    siPOU3F2

    siSMARCA5

    ***

    b

    *** *

    *

    **

    **

    LN229

    08-387 08-387

    i

    Figure 6

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/

  • Published OnlineFirst September 24, 2020.Mol Cancer Res Tiantian Cui, Erica Hlavin Bell, Joseph McElroy, et al. stemness and therapeutic response in glioblastomaA novel miR-146a-POU3F2/SMARCA5 pathway regulates

    Updated version

    10.1158/1541-7786.MCR-20-0353doi:

    Access the most recent version of this article at:

    Material

    Supplementary

    http://mcr.aacrjournals.org/content/suppl/2020/09/24/1541-7786.MCR-20-0353.DC1

    Access the most recent supplemental material at:

    Manuscript

    Authorbeen edited. Author manuscripts have been peer reviewed and accepted for publication but have not yet

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

    Subscriptions

    Reprints and

    [email protected] at

    To order reprints of this article or to subscribe to the journal, contact the AACR Publications

    Permissions

    Rightslink site. Click on "Request Permissions" which will take you to the Copyright Clearance Center's (CCC)

    .http://mcr.aacrjournals.org/content/early/2020/09/24/1541-7786.MCR-20-0353To request permission to re-use all or part of this article, use this link

    on June 8, 2021. © 2020 American Association for Cancer Research. mcr.aacrjournals.org Downloaded from

    Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited. Author Manuscript Published OnlineFirst on September 24, 2020; DOI: 10.1158/1541-7786.MCR-20-0353

    http://mcr.aacrjournals.org/lookup/doi/10.1158/1541-7786.MCR-20-0353http://mcr.aacrjournals.org/content/suppl/2020/09/24/1541-7786.MCR-20-0353.DC1http://mcr.aacrjournals.org/cgi/alertsmailto:[email protected]://mcr.aacrjournals.org/content/early/2020/09/24/1541-7786.MCR-20-0353http://mcr.aacrjournals.org/

    Article FileFigure 1Figure 2Figure 3Figure 4Figure 5Figure 6