Tirapazamine Sensitizes Hepatocellular Carcinoma Cells to Topoisomerase … · 2013. 12. 20. ·...

43
1 Tirapazamine Sensitizes Hepatocellular Carcinoma Cells to Topoisomerase I Inhibitors via Cooperative Modulation of Hypoxia-Inducible Factor-1α Tian-Yu Cai 1 , Xiao-Wen Liu 1 , Hong Zhu 1 , Ji Cao 1 , Jun Zhang 1 , Ling Ding 1 , Jian-Shu Lou 1 , Qiao-Jun He 1 *, Bo Yang 1 * 1. Zhejiang Province Key Laboratory of Anti-Cancer Drug Research, Institute of Pharmacology and Toxicology, College of Pharmaceutical Sciences, Zhejiang University, Hangzhou, 310058, China *Corresponding authors: Professor Bo Yang, Ph.D. Professor Qiao-Jun He, Ph.D. #866 Yuhangtang Road, College of Pharmaceutical Sciences, Zhejiang University, Hangzhou, 310058, China Telephone: 86-(571)88208400; Fax: 86-(571)88208400 E-mail: [email protected], [email protected] Running title: TPZ Sensitizes HCC to Topo I Inhibitors Keywords on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

Transcript of Tirapazamine Sensitizes Hepatocellular Carcinoma Cells to Topoisomerase … · 2013. 12. 20. ·...

  • 1

    Tirapazamine Sensitizes Hepatocellular Carcinoma Cells to

    Topoisomerase I Inhibitors via Cooperative Modulation of

    Hypoxia-Inducible Factor-1α

    Tian-Yu Cai1, Xiao-Wen Liu1, Hong Zhu1, Ji Cao1, Jun Zhang1, Ling Ding1, Jian-Shu

    Lou1, Qiao-Jun He1*, Bo Yang1*

    1. Zhejiang Province Key Laboratory of Anti-Cancer Drug Research, Institute of

    Pharmacology and Toxicology, College of Pharmaceutical Sciences, Zhejiang

    University, Hangzhou, 310058, China

    *Corresponding authors:

    Professor Bo Yang, Ph.D.

    Professor Qiao-Jun He, Ph.D.

    #866 Yuhangtang Road, College of Pharmaceutical Sciences, Zhejiang University,

    Hangzhou, 310058, China

    Telephone: 86-(571)88208400; Fax: 86-(571)88208400

    E-mail: [email protected], [email protected]

    Running title: TPZ Sensitizes HCC to Topo I Inhibitors

    Keywords

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 2

    Topoisomerase I Inhibitors, tirapazamine, hypoxia-inducible factor-1α, hypoxia,

    hepatocellular carcinoma

    Abbreviations

    TPZ, tirapazamine; HIF-1α, hypoxia-inducible factor-1α; PARP, poly (ADP-ribose)

    polymerase; HCC, Hepatocellular carcinoma; VEGF, vascular endothelial growth

    factor; HRE, hypoxia response elements; PI, propidium iodide; CI, Combination

    Index; DAPI, 4',6-diamidino-2-phenylindole

    .

    Grand Support

    This work was supported by grants from National Natural Science Foundation of

    China No. 81302789 (J. Lou), Zhejiang Provincial Program for The Cultivation of

    High-Level Innovative Health Talents (Q. He), the Department of Education of

    Zhejiang Province No. Y201226213 (L. Ding) and the Fundamental Research Funds

    for the Central Universities No. 2012QNA7021 (L. Ding).

    Competing Interests

    The authors disclose no potential conflicts of interest.

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 3

    Abstract

    Topoisomerase I inhibitors are a class of anti-cancer drugs with a broad spectrum of

    clinical activity. However, they have limited efficacy in hepatocellular cancer. Here,

    we present in vitro and in vivo evidence that the extremely high level of hypoxia-

    inducible factor-1α (HIF-1α) in HCC is intimately correlated with resistance to

    topoisomerase I inhibitors. In a previous study conducted by our group, we found that

    TPZ could downregulate HIF-1α expression by decreasing HIF-1α protein synthesis.

    Therefore, we hypothesized that combining TPZ with topoisomerase I inhibitors may

    overcome the chemoresistance. In this study, we investigated that in combination with

    TPZ, Topoisomerase I inhibitors exhibited synergistic cytotoxicity and induced

    significant apoptosis in several HCC cell lines. The enhanced apoptosis induced by

    TPZ plus SN-38 (the active metabolite of irinotecan) was accompanied by increased

    mitochondrial depolarization and caspase pathway activation. The combination

    treatment dramatically inhibited the accumulation of HIF-1α protein, decreased the

    HIF-1α transcriptional activation and impaired the phosphorylation of proteins

    involved in the homologous recombination repair pathway, ultimately resulting in the

    synergism of these two drugs. Moreover, the increased anticancer efficacy of TPZ

    combined with irinotecan was further validated in a human liver cancer Bel-7402

    xenograft mouse model. Taken together, our data show for the first time that HIF-1α

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 4

    is strongly correlated with resistance to topoisomerase I inhibitors in HCC. These

    results suggest that HIF-1α is a promising target and provide a rationale for clinical

    trials investigating the efficacy of the combination of topoisomerase I inhibitors and

    TPZ in hepatocellular cancers.

    Introduction

    Hepatocellular carcinoma (HCC) is the most frequently occurring primary liver

    cancer, with an incidence of half a million cases per year worldwide (1). Although

    surgery remains the treatment of choice for HCC, tumor size and hepatic functional

    reservation might restrict surgical ablation. Conventional radio/chemotherapies are

    generally not prescribed for advanced HCC due to their low efficacies (2). We and

    other groups (3, 4) have found that the expression of HIF-1α is increased in most

    HCCs and is associated with poor outcomes. Because fibrogenesis during the

    development of cirrhosis ultimately destroys the normal blood supply to the liver, the

    blood supply is limited in HCC with cirrhosis, which ultimately leads to local hypoxia

    (5). The insufficient blood supply of the rapidly growing tumor tissues also induces

    hypoxia in the central region of the tumor. Hypoxia leads to the activation of hypoxia-

    inducible factor 1 alpha (HIF-1α), which acts as the transcriptional factor for a variety

    of essential genes in the hypoxic microenvironment, including those encoding for

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 5

    vascular endothelial growth factor (VEGF) and basic fibroblast growth factor (BFGF)

    (6-8).

    HIF-1α is a master regulator of tumor growth, angiogenesis, metastasis and

    resistance to anticancer drugs (9-11). Although the proteasome pathway rapidly

    degrades HIF-1α under normoxia, this protein is stable under hypoxia, where it

    translocates to the nucleus and binds to hypoxia response elements (HREs) within the

    promoter of its target genes. Significant evidence indicates that HIF-1α plays an

    important role in the resistance to chemotherapy of HCC (12, 13). Hence, inhibition of

    the HIF-1α pathway is a promising approach for the treatment of HCC.

    Camptothecin (CPT), a natural product isolated from Camptotheca acuminata

    Decne, was originally discovered because of its antitumor activity, and the mechanism

    of action was later demonstrated to be the accumulation of Topoisomerase I–DNA

    adducts, observed both in vitro and in vivo (14-17). CPTs bind the covalent 3-

    phosphotyrosyl intermediate and specifically block DNA re-ligation, thus inducing

    DNA damage. During DNA replication, the replication fork is thought to collide with

    the ‘‘trapped’’ Topoisomerase I–DNA complexes, resulting in double-strand breaks

    and ultimately cell death. Irinotecan is a topoisomerase I inhibitor registered for the

    treatment of gastrointestinal cancer in humans and has been shown to exert activity on

    many cancer cell lines in vitro, but it fails to provide definitive information in vivo,

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 6

    particularly through intravenous infusion in HCC patients (18). Similar to other

    cancer types, HCC progresses and develops drug resistance via the upregulation of

    stress-induced molecules such as HIF-1α. We surmised that HIF-1α may play an

    important role in the resistance to topoisomerase I inhibitors in HCC.

    Tirapazamine (TPZ, 3-amino-1,2,4-benzotriazine 1,4 dioxide, SR4233) is a

    bioreductive drug belonging to the aromatic N-oxide family that has a selective

    toxicity for hypoxic cells(19). In cell suspensions, TPZ is 50- to 500-fold more toxic

    to hypoxic cells compared with cells in normoxia. The hypoxic selectivity of TPZ is

    due to the rapid reoxidation of the initial radical to the parent prodrug by O2. Thus,

    TPZ functions as a hypoxia-activated agent (20, 21), exhibiting the capability of

    potentiating the antitumor efficacy of many anticancer drugs (22-25). The putative

    mechanism may be attributed to its suppression of the accumulation of HIF-1α protein.

    Thus, we hypothesized that the combination of TPZ (HIF-1α–downregulating agents)

    with topoisomerase I inhibitors may be a potential therapeutic regimen for HCC.

    In our study, we show for the first time that HIF-1α is strongly correlated with

    resistance to topoisomerase I inhibitors in HCC. Using both in vitro and in vivo

    models of HCC, we demonstrate the potential for synergistic activity of TPZ in

    combination with topoisomerase I inhibitors based on the modulation of the HIF-1α

    protein.

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 7

    Materials and Methods

    Reagents and cell lines

    CPT-11, SN-38, TPT, HCPT and MONCPT were kindly provided by Dr. Wei Lu (East

    China Normal University, Shanghai, China), and TPZ (Tirapazamine) was purchased

    from Topharman Shanghai Co. Ltd. All of the drugs were dissolved in

    dimethylsulfoxide (DMSO) and stored in aliquots (20 mM) at -20 °C. Primary

    antibodies to cleaved-Caspase 3, H2AX, γ-H2AX, Chk1, p-Chk1, Chk2, p-Chk2,

    ATM, p-ATM, Rad51 and β-Actin were obtained from Cell Signaling Technologies.

    Antibodies to Caspase-3, PARP, and HRP-labeled secondary anti-goat, anti-mouse,

    and anti-rabbit antibodies were from Santa Cruz Biotechnology. Antibody to HIF-1α

    was obtained from BD Biosciences. DMSO, propidium iodide (PI), sulforhodamine B

    (SRB), DAPI (4', 6-diamidino-2-phenylindole), JC-1 and CoCl2 were obtained from

    Sigma-Aldrich.

    The five HCC cell lines, HepG2, Hep3B, Huh 7, Bel-7402 and SMMC-7721 were

    purchased from Cell Bank of China Science. A normal liver cell line, Chang liver, was

    obtained from Cell Bank of China Science. Following receipt, cells were grown and

    frozen as a seed stock as they were available. All six cell lines were passaged for a

    maximum of 2 months, after which new seed stocks were thawed. All six cell lines

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 8

    were authenticated using DNA fingerprinting (variable number of tandem repeats),

    confirming that no cross-contamination occurred during this study. All six cell lines

    were tested for mycoplasma contamination at least every month. The six cell lines

    were cultured in RPMI 1640 medium/DMEM containing 10% fetal bovine serum and

    1% antibiotics in a humidified atmosphere with 5% CO2 at 37 °C. Hypoxia treatment

    was carried out by placing the cells in a CO2-Water Jacketed Incubator (Thermo

    Forma) filled with a mixture of 0.6% O2, 5% CO2, and 94.4% nitrogen.

    Cell proliferation assay

    Cell proliferation was evaluated by the Sulforhodamine B (SRB) protein staining

    assay (26). Briefly, treated cells in 96-well plates were fixed with 10% TCA and

    stained with SRB. SRB in the cells was dissolved in 10 mM Tris-HCl and measured at

    515 nm using a multiwell spectrophotometer (Thermo Electron Co.). The rate of

    inhibition of cell proliferation was calculated for each well as (A515control cells-

    A515treated cells)/A515control cells×100%. The average IC50 values were determined by the

    logit method from at least three independent tests.

    Flow cytometric analysis of apoptosis and mitochondrial membrane

    potential (ΔΨm)

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 9

    Cells were treated with SN-38, TPZ or their combination under normoxia or hypoxia

    for 12 h. After harvesting and washing twice with cold PBS buffer, the Annexin V-

    FITC/PI apoptosis Detection Kit was used to analyze apoptotic cells. Analysis of the

    sub-G1 phase after PI staining was also employed to assess apoptosis. For PI staining,

    treated cells were harvested and fixed with 70% ethanol at -20 oC, then incubated with

    RNaseA followed by PI in the dark for 30 min. For determination of mitochondrial

    potential, cells were resuspended in PBS containing 0.1 µΜ JC-1 and were incubated

    at 37 °C for 15 min in the dark. All samples were analyzed using a FACS-Calibur

    cytometer (Becton Dickinson).

    Clonogenic assays

    Cells treated with TPZ or SN-38/TPT/HCPT/MONCPT or their combination were

    plated in 60-mm dishes in triplicate on soft agar. Once set, the dishes were overlaid

    with 2.5 ml of medium and incubated at 37 °C for 10 days in hypoxic conditions, at

    which time the colonies were scored and photographed.

    Quantitative real time RT–PCR (qRT–PCR) analysis

    Total RNA was prepared with Trizol (Invitrogen) and cDNA was synthesized from 2

    μg of RNA with Revertra Ace (Toyobo). qRT–PCR was performed with SYBR Green

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 10

    PCR kits (Qiagen Inc., Valencia, CA). The primers used were as follows: HIF-1α,

    forward primer 5′-TCACCACAGGACAGTACAGGATGC-3′ and reverse primer 5′-

    CCAGCAAAGTTAAAGCATCAGGTTCC-3′; VEGF, forward primer 5′-

    AGGAGGGCAGAATCATCACG-3′ and reverse primer 5′-

    CAAGGCCCACAGGGATTTTCT-3′; GAPDH, forward primer 5′-

    GTCATCCATGACAACTTTGG-3′ and reverse primer 5′-

    GAGCTTGACAAAGTGGTCGT-3′. In all experiments, two negative controls were

    also performed. Values are expressed as fold increases relative to the reference sample.

    Immunofluorescence

    Cells were plated onto glass culture slides and incubated with SN-38/TPZ or SN-38 +

    TPZ or vehicle [0.1% DMSO (v/v)] for different time periods. Cells were then fixed

    with 4% paraformaldehyde and permeabilized with PBS containing 0.1% Triton X-

    100. After blocking with 5% bovine serum albumin for 30 min, cells were incubated

    with primary HIF-1α or γ-H2AX antibodies (1:100 dilution) for 1 h, washed 3x with

    PBS and then incubated with Alexa Fluor 488–conjugated and rhodamine secondary

    antibodies (Invitrogen) respectively in the dark. Nuclei were visualized by staining

    with DAPI (4′6-diamidino-2-phenylindole). Fluorescence signals were analyzed using

    an Olympus Fluorview 1000 confocal microscope.

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 11

    Plasmid based overexpression of HIF-1α

    Transfection with HIF-1α plasmids or empty vector (EV) obtained form Origene was

    performed using the FuGENE® HD Transfection Reagent (Roche Applied Science)

    according to the manufacturer’s instructions.

    shRNA knockdown of HIF-1α

    shRNA targeting HIF-1α was purchased from Origene with a GFP-tag and puromycin

    resistance. Transfection of shRNA into Bel-7402 cells was performed using

    Lipofectamine 2000 (Invitrogen). Individual clones stably expressing the shRNA were

    generated by single-cell sorting into 96 round-bottom wells and antibiotic selection

    with puromycin (500 ng/mL). Western blot analysis and an Olympus Fluorview 1000

    confocal microscope were used to select clones with stable knockdown of HIF-1α.

    Immunoblot analysis

    Immunoblotting of biomarkers in cell lysates and tumor tissue homogenates was

    performed as previously described(27).

    Luciferase reporter assays

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 12

    Bel-7402 cells were seeded into 96-well plates and then cotransfected with HRE-

    luciferase-pGL3 and renilla luciferase (internal control for transfection efficiency)

    plasmids using Lipofectamine 2000 according to the manufacturer's instructions.

    Luciferase activity was measured using the Dual-Luciferase reporter assay system. In

    the assay, firefly luciferase activity was normalized by renilla luciferase.

    Measurement of in vivo activity

    Tumors were established by injection of Bel-7402 cells (5×106 cells per animal,

    subcutaneously into the armpit) into 5-to 6-week-old BALB/c male athymic mice.

    Treatments were initiated when tumors reached a mean group size of about 100 mm3.

    Tumor volume (mm3) was measured with calipers and calculated as (W2×L)/2, where

    W is the width and L is the length. Athymic mice were intraperitoneally injected with

    CPT-11 (2.5 mg/kg) dissolved in physiological saline once daily and TPZ (25 mg/kg)

    dissolved in a cremophor:ethanol:0.9% sterile sodium chloride solution (1:1:8,

    volume) every two days. Mouse weight and tumor volumes were recorded every 2

    days until the animals were sacrificed. Animal care was in accordance with

    institutional guidelines.

    Statistical analyses

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 13

    The results are expressed as the mean ± SD of at least three independent experiments.

    Differences between means were analyzed using Student’s t-test and were considered

    statistically significant when P < 0.05.

    Results

    HIF-1α protein level is strongly correlated with resistance to

    topoisomerase I inhibitors in HCC cell lines

    In our study, we first investigated the strength of the association between

    expression of HIF-1α and resistance to topoisomerase I inhibitors in HCC cell lines.

    We measured Bel-7402 cell growth under hypoxic or normoxic culturing conditions in

    the presence of topoisomerase I inhibitors. Compared to normoxia, the IC50s for SN-

    38, TPT, HCPT and MONCPT were 7.6-, 6.2-, 5.4- and 4.3-fold higher, respectively,

    under 0.6% O2. Then, we decided to investigate whether the accumulation of HIF-1α

    induced by hypoxia was a key factor in this resistance; thus, an HIF-1α-deficient Bel-

    7402 cell line was established. The transfection of a HIF-1α-targeting shRNA

    construct almost completely suppressed the protein expression of HIF-1α even under

    hypoxia, while HIF-1α expression in empty-vector-transfected Bel-7402 cells

    remained intact (Fig. 1A). Compared to normoxia, the IC50s for SN-38, TPT, HCPT

    and MONCPT were similar under 0.6% O2 for Bel-7402/HIF-1α (-), hypoxia-induced

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 14

    resistance disappeared in HIF-1α gene silencing cells (Fig. 1B and Fig. S1).

    Meanwhile, cells treated with 200 µM CoCl2, which resulted in a robust induction of

    intracellular HIF-1α, showed no statistical difference in the IC50 for SN-38 between

    cells treated with CoCl2 and hypoxic cells (Fig. 1C). These data demonstrate that

    hypoxia can induce resistance to topoisomerase I inhibitors in HCC cell lines and

    HIF-1α may play an important role.

    Tirapazamine synergistically increases topoisomerase I inhibitor-

    inhibition of cell proliferation in HCC cell line

    The above observations indicate the therapeutic potential treatment with the

    combination of topoisomerase I inhibitors and HIF-1α inhibitors, which may

    overcome the resistance to topoisomerase I inhibitors in HCC cell lines. It has been

    reported that TPZ can inhibit the accumulation of HIF-1α protein (28), and we

    confirmed this effect in our system using luciferase reporter gene assays and

    immunoblot analysis. As shown in Fig. 2B, luciferase activity was markedly increased

    under hypoxia compared with normoxia, while this activity was suppressed by TPZ in

    a dose-dependent manner after 6 h treatment, ranging from 10.2% (0.31 μM) to

    66.2% (10 μM). Similarly, the protein level of HIF-1α was reduced upon TPZ

    treatment (Fig. 2A). Taken together, these data imply that TPZ is indeed inhibiting

    HIF-1α accumulation.

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 15

    Because TPZ can inhibit the accumulation of HIF-1α protein, we determined the

    cytotoxicity of TPZ and each of the four topoisomerase I inhibitors alone and of TPZ

    combined with each of the other drugs at clinically achievable concentrations in the

    Bel-7402 cell line using the SRB cytotoxicity assay. CI values were calculated at the

    fixed-ratio drug concentrations used for cytotoxicity assays. Dose-response curves to

    TPZ, the four topoisomerase I inhibitors and TPZ combined with each topoisomerase

    I inhibitor are shown in Fig. 2C-F. The combinations of TPZ plus topoisomerase I

    inhibitor all showed strong synergy (CI < 0.3) or synergy (CI < 0.7) in the Bel-7402

    cell line, achieving about 80% cell death. In addition, we determined the cytotoxicity

    of TPZ (2.5 μM) in combination with various concentrations of topoisomerase I

    inhibitors in Bel-7402 cells. As shown in Fig. S2, treatment of cells with TPZ (2.5 μM)

    significantly enhanced cell sensitivity to topoisomerase I inhibitors. To analyze the

    antitumor effect of the combination treatment of TPZ with the four topoisomerase I

    inhibitors, a clonogenic assay was performed. As shown in Fig. S3, TPZ and the

    topoisomerase I inhibitors both weakly suppressed colony formation of Bel-7402

    cancer cells (55–85%). Significantly, TPZ enhanced the anti-tumor effect of

    topoisomerase I inhibitors and drastically suppressed colony formation (11–23%).

    Thus, the combination effect of TPZ and topoisomerase I inhibitors was quite

    dramatic, and this effect had to be confirmed in other HCC cell lines.

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 16

    Tirapazamine synergistically increases SN-38-induced cell proliferation

    inhibition in different HCC cell lines

    Because TPZ synergized most strongly with SN-38 (active metabolite of

    irinotecan), and irinotecan is a widely used drug in clinical practice, we determined

    the effects of TPZ, SN-38 and their combination on various HCC cell lines, including

    Bel-7402, SMMC-7721, HepG2, Hep3B and Huh-7. Survival curves to TPZ, SN-38

    and their combination are shown in Fig. 3A-E. As shown in Table S1, SN-38 plus

    TPZ showed distinct synergy in all cancer cell lines tested, with all CI values below

    0.8, indicating the much stronger antiproliferative abilities achieved by the

    combination of SN-38 and TPZ. However, in a normal liver cell line, the combination

    effect was minimal (Fig. 3F).

    TPZ sensitizes HCC cells to SN-38-triggered caspase-dependent

    apoptosis

    To further understand mechanisms of enhanced cytotoxicity observed with the

    combination of TPZ and SN-38, we examined their effects on apoptosis,

    mitochondrial membrane depolarization, and caspase activation. We first investigated

    whether the observed cytotoxicity was due to apoptosis. DAPI staining was used to

    visualize the apoptosis induced by the co-treatment of TPZ and SN-38. After exposure

    to SN-38 or/and TPZ of the indicated concentrations for 12 h, the nuclear DNA of

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 17

    Bel-7402 cells were permeabilized and stained with DAPI. As shown in Fig. 4A, 10

    µM TPZ plus 0.1 µM SN-38 triggered more apoptosis, as indicated by the apoptotic

    bodies, than the mono-treated cells. The PI (sub-G1) staining assay and Annexin V-PI

    staining assay were also employed to detect levels of apoptosis (Fig. 4B and 4D).

    After treatment with TPZ and/or SN-38 for 12 h, we found that cells treated with the

    two agents together experienced 49% and 33% apoptosis in Bel-7402 cells using the

    two assays. This was more apoptosis than either TPZ or SN-38 induced alone. We

    next investigated the effects of treatment with the single agents or the combination on

    mitochondrial membrane potential. Mitochondrial membrane depolarization, as

    determined by the JC-1 mitochondrial probe, was induced by TPZ and SN-38 as

    single agents 12 hours after treatment, and the combination resulted in a greater than

    additive effect (P < 0.05) compared with the single agents (Fig. 4C).

    Furthermore, we examined the effect of TPZ, SN-38 and their combination on the

    activation of caspase-3 and cleavage of poly (ADP-ribose) polymerase (PARP). We

    found that although TPZ and SN-38 had little effect on caspase-3 and PARP, the two

    together induced a more significant cleavage of PARP and caspase-3 (Fig. 4E).

    Together, our results demonstrate that the combination of TPZ and SN-38 elicited

    more apoptosis via mitochondrial membrane depolarization, resulting in increased cell

    death.

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 18

    TPZ and SN-38 combination therapy suppresses HIF-1α expression

    Our results indicate that high levels of HIF-1α confer resistance to SN-38; thus, we

    were interested in examining the involvement of HIF-1α in the efficacy of the TPZ

    and SN-38 combination treatment. Western blot analysis and immunofluorescence

    microscopy (Fig. 5A and 5B) showed that TPZ and SN-38 alone weakly inhibited the

    accumulation of HIF-1α protein in vitro. Strikingly, combination therapy resulted in

    complete inhibition of HIF-1α protein accumulation. The combination did not have an

    effect on the HIF-1β and HIF-2α proteins. Then, we analyzed the mRNA levels of

    HIF-1α and VEGF by RT-PCR. The results showed that exposure to TPZ or SN-38

    reduced the expression of VEGF, a target gene of HIF-1α (Fig. 5C). This effect was at

    least additive when the two agents were administered together, similar to their effect

    on HIF-1α protein levels. HIF-1α binds to the HRE cis-elements within the promoters

    of hypoxia-responsive genes to regulate their expression. To investigate whether the

    TPZ and SN-38 combination treatment could affect the activity of HIF-1α, we

    examined its effect on HIF-1α binding to DNA using a reporter gene assay. After

    transfection, cells were exposed to the single agents or the combination for 4 h under

    hypoxia. As shown in Fig. 5D, luciferase activity was markedly induced under

    hypoxia compared with normoxia, and this activity was both inhibited by increasing

    doses of TPZ or SN-38. This inhibition was dramatically increased when TPZ and

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 19

    SN-38 were combined, and at the maximum combination dose (10 µM plus 0.1 µM)

    there was a 92% decrease of HIF-1α transcriptional activity.

    Next, we investigated the effect of the combination on HIF-1α posttranscriptional

    regulation. As shown in Fig. S4A, the degradation rates of HIF-1α were similar in

    treated and untreated cells, indicating the combination did not significantly affect

    HIF-1α degradation. To exclude the possibility that the inhibition of HIF-1α

    accumulation was attributed to HIF-1α protein degradation, Bel-7402 cells were

    pretreated with MG132 (a specific proteasome inhibitor) or chloroquine (CQ, a

    lysosome inhibitor) before the drugs were added to prevent HIF-1α degradation. As

    shown in Fig. S4B, MG132 and CQ treatment failed to reverse the decrease of HIF-

    1α protein levels in the combination treatment.

    To more directly assess the effects of the combination on the rate of de novo

    synthesis of the HIF-1α protein, cells were pretreated with cycloheximide (CHX) for

    three hours under normoxia to inhibit new protein synthesis and then incubated in

    fresh medium. At this point, most HIF-1α proteins were degraded, and a limited

    amount of protein remained. These CHX-pretreated cells were then exposed to

    hypoxic conditions, followed by treatment with TPZ, SN-38 or the combination for

    different periods of time, and HIF-1α protein levels were analyzed by western blot.

    As shown in Fig. 5E, significantly less HIF-1α protein accumulated in cells treated

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 20

    with the combination than in cells treated with TPZ or SN-38 alone at all times tested.

    This result further confirmed that the combination decreases the rate of HIF-1α

    protein synthesis.

    There is a functional link between HIF-1α variation and DNA repair. To assess the

    effect of TPZ, SN-38 and their combination on DNA, the formation of DSBs was

    measured by immunostaining of γ-H2AX. The cells treated with TPZ or SN-38

    showed several γ-H2AX foci after a 4-hour treatment. Cells treated with the

    combination treatment for 4 hours showed a greater than 3-fold increase in the

    number of γ-H2AX foci (Fig. 5B), which is similar to the results for γ-H2AX protein

    levels (Fig. 5A). To investigate whether the effect of the combination treatment on

    HIF-1α expression might affect DNA repair, shRNA-mediated repression of HIF-1α

    was performed in Bel-7402 cell line. Bel-7402/HIF-1α(+) and Bel-7402/HIF-1α(-)

    cells were pretreated with SN-38 for two hours under hypoxia to form DSBs and then

    incubated in fresh medium without drug for six hours. The formation of DSBs was

    measured by immunostaining. As shown in Fig. 5F, the Bel-7402/HIF-1α(-) cells

    contained more DSB-indicative γ-H2AX than those Bel-7402/HIF-1α(+) cells,

    indicating HIF-1α knockdown causes an inhibition of DNA repair, resulting in a larger

    population of persistently unrepaired DSBs.

    CPT-induced DSB are predominantly repaired by homologous recombination (HR)

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 21

    in mammalian cells (29-31). We attempted to detect which repair pathway was

    inhibited by HIF-1α in the current model to repair the DSBs elicited by CPT. As

    shown in Fig. 5A, ATM, Chk1 and Chk2 were phosphorylated after a 4 h of exposure

    to SN-38 or TPZ. The Rad51 and phosphorylation of Chk1 at Ser317 were both

    decreased in the combination treatment of TPZ and SN-38. The close correlation

    between loss of RAD51 and DSB accumulation suggested that the inhibition of the

    RAD51-mediated HR repair pathway was likely responsible for the accumulation of

    DSBs in Bel-7402 cells.

    The combination of TPZ and irinotecan arrests tumor growth.

    Because the TPZ plus SN-38 exhibited the most effective synergistic effect with the

    lowest CI values on human HCC Bel-7402 cells (Table S1), the in vivo efficacy of the

    TPZ and irinotecan combination therapy was chosen to test against Bel-7402

    xenografts in nude mice. As shown in Fig. 6A-B and Table S2, the i.p. administration

    of TPZ at a dose of 25 mg/kg every two days or irinotecan at a dose of 2.5 mg/kg

    every day produced no significant difference in mean RTV compared with that of the

    control group. However, TPZ plus irinotecan caused marked tumor growth inhibition

    (T/C value: 36.9%) that was significantly greater than that caused by TPZ (T/C value:

    88.1%) or irinotecan treatment alone (T/C value: 86.4%). Furthermore, compared

    with the initial body weights, the mice treated with the combination showed no

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 22

    significant body weight loss on day 26 (Fig. 6C). Thus, the combination of TPZ and

    irinotecan exerted more potent tumor growth inhibitory effects compared with the

    monotreatment groups, but caused no extended bodyweight loss to the animals.

    The combination therapy induces apoptosis and downregulates HIF-1α

    in tumor tissues.

    Hematoxylin & eosin (H&E) staining for formalin-fixed paraffin embedded tissues

    was used to distinguish tumor tissues from adjacent normal tissues. In mice treated

    with TPZ and irinotecan, most tumor cells were severely damaged or destroyed (Fig.

    6E). We next aimed to explore the effect of mono-treatment or combined treatment of

    TPZ and irinotecan on the protein levels of apoptosis-related proteins in tumor tissues

    from drug-administrated mice. Notably, the caspase cascade was activated in the

    tissues from the nude mice treated with combination therapy (Fig. 6D). Importantly,

    the expression of HIF-1α, as determined by western blot (Fig. 6D) and

    immunohistochemistry (Fig. 6F), was consistent with the aforementioned cell culture

    data (Fig. 5A and 5B), highlighting the involvement of these proteins in the tumor

    growth inhibitory effects exerted by TPZ and irinotecan in vivo.

    Discussion

    Topoisomerase I inhibitors are a class of anti-cancer agents with a mechanism of

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 23

    action involving the disruption of DNA replication in cancer cells, the result of which

    is cell death. They are important components in the current treatment of colorectal

    cancer; however, in clinical studies of topoisomerase I inhibitors with advanced

    hepatocellular carcinoma, they had modest activity in advanced hepatocellular cancer.

    Accumulating evidence demonstrates that overexpression of HIF-1α plays an

    important role in chemoresistance and that hypoxia has a protective effect against

    DNA damage induced by topoisomerase inhibitors (32, 33). As shown in Fig. S5A,

    the protein level of HIF-1α in Bel-7402 cells in hypoxia is much higher than the levels

    in two colon cancer cell lines HT29 and SW480, which were reported to be

    susceptible to SN-38 (34) in hypoxia owing to their decreased levels of HIF-1α. Thus,

    we speculated that the HCC resistance to topoisomerase I inhibitors is correlated with

    levels of HIF-1α. As expected, under hypoxia, Bel-7402 cells were resistant to SN-38,

    TPT, HCPT and MONCPT. Furthermore, our results using short hairpin RNA or

    cobalt chloride (CoCl2) are also consistent with our hypothesis that HIF-1α is an

    important element in the limited efficacy of topoisomerase I inhibitors on HCC.

    Accordingly, selective down-regulation of HIF-1α has been documented to be a

    mechanism underlying increased sensitization to topoisomerase I inhibitors.

    Tirapazamine is an anticancer drug that is activated to a toxic radical specifically in

    hypoxic microenvironments. Cells in hypoxic areas are resistant to killing by most

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 24

    anticancer drugs, and TPZ has been shown to potentiate the antitumor efficacy of

    many anticancer drugs in hypoxia. It was reported that TPZ can induce a reduction in

    HIF-1α protein levels; thus the combination of TPZ with topoisomerase I inhibitors

    was of particular interest to us.

    As expected, in our in vitro study, synergistic anticancer effects of TPZ plus

    topoisomerase I inhibitors were observed in human HCC cells. CI values and the

    significant decline of the survival curves in the combination group strongly

    demonstrated that TPZ potentiated the SN-38-induced cytotoxicity in HCC cells. In

    addition, the data from both the flow cytometry and western blot analyses indicated

    that TPZ plus SN-38 synergistically increased the execution of caspase-dependent

    apoptosis. We also compared the cytotoxicity of TPZ plus SN-38 in HCC cell lines to

    the Chang liver cell line (normal liver cells). The data suggest that the combination of

    TPZ plus SN-38 kills HCC cells efficiently but minimally affects normal liver cells.

    In our in vivo experiment, this synergistic effect of the drug combination was also

    observed in the Bel-7402 xenograft nude mice model (Fig. 6B). The co-administration

    of TPZ and irinotecan arrested tumor growth by 60.9%. Moreover, there was no

    significant difference in body weight loss between combination and irinotecan

    treatment groups. These results suggest that the TPZ and irinotecan combination

    synergistically inhibited tumor growth and had minimal toxicity in vivo.

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 25

    In this work, we have shown that the combination of TPZ and topoisomerase I

    inhibitors has a major antitumor effect in vivo and induces massive cell death in vitro

    under hypoxic conditions. These effects are correlated with a potent inhibition of HIF-

    1α accumulation in tumor cells. Adding TPZ to SN-38-based therapy was

    accompanied by inhibition of HIF-1α accumulation in vivo and in vitro as well as a

    dramatic reduction of tumor volume and massive tumor cell death. This effect was

    observed even at low doses of TPZ and SN-38 due to a cooperative effect of the two

    agents on HIF-1α expression. The combination of TPZ and SN-38 did not affect the

    mRNA levels of HIF-1α but did decrease the rate of HIF-1α protein synthesis and

    inhibited HIF-1α transcriptional activation and the accumulation of HIF-1α protein,

    resulting in a decrease in expression of target genes, such as VEGF.

    The antitumor activity of topoisomerase I inhibitors is generally believed to result

    from stabilization of covalent topoisomerase I-DNA complexes (35), which are

    converted to DNA double-strand breaks upon collision with the replication fork. TPZ,

    a hypoxia-selective cytotoxin, has demonstrated activity in cancer clinical trials.

    Under hypoxic conditions, TPZ is reduced to a topoisomerase IIα poison that leads to

    DNA double-strand breaks, single-strand breaks, and base damage (20). The typical

    mechanism by which both TPZ and topoisomerase I inhibitors work is through

    binding to the topoisomerase molecule, which blocks the topoisomerase from binding

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 26

    to the DNA, causing the DNA damage and ultimately leading to cell death.

    HIF-1α plays an important role in the prevention and repair of drug-induced DNA

    damage and can cause chemoresistance. In Minori Koshiji’s work (36), transcript

    levels of two DNA-PK complex members, DNA-PKcs and Ku80, were found to be

    downregulated in HIF-1α-deficient cells that were previously found to have a higher

    susceptibility to DNA DSB-inducing chemotherapeutics. L E Huang reported (37) that

    HIF-1α upregulates DNA repair genes by counteracting c-Myc activities through c-

    Myc displacement. In recent research (38), modulating expression of HIF-1α by small

    interfering RNA or cobalt chloride markedly reduced or increased transcription of

    XPA in lung cancer cell lines, and XPA has a dual role in sensing and recruiting other

    DNA repair proteins to the damaged template. Therefore, we hypothesized that

    modulation on HIF-1α is involved in the synergistic effect of the TPZ and SN-38

    combination treatment. To test this hypothesis, we next asked whether the synergism

    could be achieved in cells with forced HIF-1α over-expression. Bel-7402 cells forced

    to overexpress HIF-1α were treated with TPZ, SN-38 or their combination. We found

    that the synergistic effects were weakened when cells were transfected with the HIF-

    1α plasmid (Fig. S6).

    In our research, the well-established DNA damage marker γ-H2AX was used to

    investigate genetic instability. Due to the presence of HIF-1α, TPZ or SN-38 by itself

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 27

    in a low dose did not induce much DNA damage, but the combination treatment of

    TPZ and SN-38 greatly increased the amount DNA damage observed in the cells. The

    complete inhibition of HIF-1α accumulation may be responsible for γ-H2AX

    induction in our model because HIF-1α can induce a DNA damage repair response.

    Furthermore, the first phase occurred in both Bel-7402/HIF-1α (-) and Bel-7402/HIF-

    1α (+) cells after 4 hours of SN-38 treatment, which represented a fast increase in foci.

    This rapid increase in γ-H2AX foci suggests DNA damage is occurring in the tumor

    cells. In the second phase, the γ-H2AX foci levels dramatically decreased in Bel-

    7402/HIF-1α (+) cells but remained the same in Bel-7402/HIF-1α (-) cells. These data

    may be explained by the following mechanistic model: the addition of SN-38 induced

    DNA DSBs; DNA repair mechanisms were ongoing, which eliminated the increase of

    foci production; after the cells had been placed in fresh medium without SN-38 for 4h,

    the γ-H2AX foci were eliminated; the knockdown of HIF-1α by shRNA impaired

    DNA repair, a larger population of persistently unrepaired DSBs remained. We also

    observed that ATM, Chk1 and Chk2 were phosphorylated after 4 h exposure to SN-38

    or TPZ. However, Rad51 and phosphorylation of Chk1 at Ser317 were decreased in

    the combination treatment of TPZ and SN-38. The strong correlation between RAD51

    removal and DSB accumulation suggested that the inhibition of the RAD51-mediated

    HR repair pathway was most likely responsible for the accumulation of DSBs in Bel-

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 28

    7402 cells. Therefore, our observations suggested that through cooperative inhibition

    of HIF-1α accumulation, the combination of TPZ and SN-38 inhibited the RAD51-

    mediated homologous recombination repair pathway, induced the accumulation of

    topoisomerase inhibitor-induced DNA damage, exhibited synergistic cytotoxicity and

    triggered significant apoptosis in HCC cell lines.

    In summary, we report for the first time that HIF-1α is strongly correlated with

    HCC cell line resistance to topoisomerase I inhibitors. The combination significantly

    improved the anticancer activities of the drugs, as indicated by the synergistic

    inhibitory effects on cancer cell proliferation, the sensitized execution of apoptosis,

    and the enhanced in vivo antitumor efficiency. We believe these results can be

    attributed to the downregulation of HIF-1α. Collectively, these data favor the regimen

    of combining TPZ and topoisomerase I inhibitors as a promising therapeutic strategy,

    and further preclinical and clinical studies of this novel combination are warranted.

    References

    1. Mendez-Sanchez N, Villa AR, Vazquez-Elizondo G, Ponciano-Rodriguez G, Uribe M. Mortality

    trends for liver cancer in Mexico from 2000 to 2006. Ann Hepatol. 2008;7:226-9.

    2. Cheung PF, Cheng CK, Wong NC, Ho JC, Yip CW, Lui VC, et al. Granulin-epithelin precursor is

    an oncofetal protein defining hepatic cancer stem cells. PLoS One. 2011;6:e28246.

    3. Liang Y, Zheng T, Song R, Wang J, Yin D, Wang L, et al. Hypoxia-mediated sorafenib resistance

    can be overcome by EF24 through Von Hippel-Lindau tumor suppressor-dependent HIF-1alpha

    inhibition in hepatocellular carcinoma. Hepatology. 2013;57:1847-57.

    4. Zhang Q, Bai X, Chen W, Ma T, Hu Q, Liang C, et al. Wnt/beta-catenin signaling enhances

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 29

    hypoxia-induced epithelial-mesenchymal transition in hepatocellular carcinoma via crosstalk with hif-

    1alpha signaling. Carcinogenesis. 2013.

    5. Rosmorduc O, Housset C. Hypoxia: a link between fibrogenesis, angiogenesis, and

    carcinogenesis in liver disease. Semin Liver Dis. 2010;30:258-70.

    6. Bhaskar A, Gupta R, Sreenivas V, Rani L, Kumar L, Sharma A, et al. Synergistic effect of

    vascular endothelial growth factor and angiopoietin-2 on progression free survival in multiple

    myeloma. Leuk Res. 2013;37:410-5.

    7. Tartour E, Pere H, Maillere B, Terme M, Merillon N, Taieb J, et al. Angiogenesis and immunity:

    a bidirectional link potentially relevant for the monitoring of antiangiogenic therapy and the

    development of novel therapeutic combination with immunotherapy. Cancer Metastasis Rev.

    2011;30:83-95.

    8. Shin JM, Kim J, Kim HE, Lee MJ, Lee KI, Yoo EG, et al. Enhancement of differentiation

    efficiency of hESCs into vascular lineage cells in hypoxia via a paracrine mechanism. Stem Cell Res.

    2011;7:173-85.

    9. Yeligar SM, Machida K, Tsukamoto H, Kalra VK. Ethanol augments RANTES/CCL5 expression in

    rat liver sinusoidal endothelial cells and human endothelial cells via activation of NF-kappa B, HIF-1

    alpha, and AP-1. J Immunol. 2009;183:5964-76.

    10. Dolt KS, Mishra MK, Karar J, Baig MA, Ahmed Z, Pasha MA. cDNA cloning, gene organization

    and variant specific expression of HIF-1 alpha in high altitude yak (Bos grunniens). Gene. 2007;386:73-

    80.

    11. May D, Itin A, Gal O, Kalinski H, Feinstein E, Keshet E. Ero1-L alpha plays a key role in a HIF-1-

    mediated pathway to improve disulfide bond formation and VEGF secretion under hypoxia:

    implication for cancer. Oncogene. 2005;24:1011-20.

    12. Kim Y, Jang M, Lim S, Won H, Yoon KS, Park JH, et al. Role of cyclophilin B in tumorigenesis

    and cisplatin resistance in hepatocellular carcinoma in humans. Hepatology. 2011;54:1661-78.

    13. Tak E, Lee S, Lee J, Rashid MA, Kim YW, Park JH, et al. Human carbonyl reductase 1

    upregulated by hypoxia renders resistance to apoptosis in hepatocellular carcinoma cells. J Hepatol.

    2011;54:328-39.

    14. Pommier Y. Topoisomerase I inhibitors: camptothecins and beyond. Nat Rev Cancer.

    2006;6:789-802.

    15. Pommier Y. DNA topoisomerase I inhibitors: chemistry, biology, and interfacial inhibition.

    Chem Rev. 2009;109:2894-902.

    16. Pommier Y, Leo E, Zhang H, Marchand C. DNA topoisomerases and their poisoning by

    anticancer and antibacterial drugs. Chem Biol. 2010;17:421-33.

    17. Staker BL, Hjerrild K, Feese MD, Behnke CA, Burgin AB, Jr., Stewart L. The mechanism of

    topoisomerase I poisoning by a camptothecin analog. Proc Natl Acad Sci U S A. 2002;99:15387-92.

    18. Boige V, Taieb J, Hebbar M, Malka D, Debaere T, Hannoun L, et al. Irinotecan as first-line

    chemotherapy in patients with advanced hepatocellular carcinoma: a multicenter phase II study with

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 30

    dose adjustment according to baseline serum bilirubin level. Eur J Cancer. 2006;42:456-9.

    19. Sonoda A, Nitta N, Ohta S, Nitta-Seko A, Nagatani Y, Takahashi M, et al. Enhanced antitumor

    effect of tirapazamine delivered intraperitoneally to VX2 liver tumor-bearing rabbits subjected to

    transarterial hepatic embolization. Cardiovasc Intervent Radiol. 2011;34:1272-7.

    20. Peters KB, Brown JM. Tirapazamine: a hypoxia-activated topoisomerase II poison. Cancer Res.

    2002;62:5248-53.

    21. Evans JW, Chernikova SB, Kachnic LA, Banath JP, Sordet O, Delahoussaye YM, et al.

    Homologous recombination is the principal pathway for the repair of DNA damage induced by

    tirapazamine in mammalian cells. Cancer Res. 2008;68:257-65.

    22. Rischin D, Narayan K, Oza AM, Mileshkin L, Bernshaw D, Choi J, et al. Phase 1 study of

    tirapazamine in combination with radiation and weekly cisplatin in patients with locally advanced

    cervical cancer. Int J Gynecol Cancer. 2010;20:827-33.

    23. Emmenegger U, Morton GC, Francia G, Shaked Y, Franco M, Weinerman A, et al. Low-dose

    metronomic daily cyclophosphamide and weekly tirapazamine: a well-tolerated combination regimen

    with enhanced efficacy that exploits tumor hypoxia. Cancer Res. 2006;66:1664-74.

    24. Reck M, von Pawel J, Nimmermann C, Groth G, Gatzemeier U. [Phase II-trial of tirapazamine

    in combination with cisplatin and gemcitabine in patients with advanced non-small-cell-lung-cancer

    (NSCLC)]. Pneumologie. 2004;58:845-9.

    25. Johnson CA, Kilpatrick D, von Roemeling R, Langer C, Graham MA, Greenslade D, et al. Phase I

    trial of tirapazamine in combination with cisplatin in a single dose every 3 weeks in patients with solid

    tumors. J Clin Oncol. 1997;15:773-80.

    26. Papazisis KT, Geromichalos GD, Dimitriadis KA, Kortsaris AH. Optimization of the

    sulforhodamine B colorimetric assay. J Immunol Methods. 1997;208:151-8.

    27. Shin DM, Zhang H, Saba NF, Chen AY, Nannapaneni S, Amin AR, et al. Chemoprevention of

    head and neck cancer by simultaneous blocking of epidermal growth factor receptor and

    cyclooxygenase-2 signaling pathways: preclinical and clinical studies. Clin Cancer Res. 2013;19:1244-56.

    28. Zhang J, Cao J, Weng Q, Wu R, Yan Y, Jing H, et al. Suppression of hypoxia-inducible factor

    1alpha (HIF-1alpha) by tirapazamine is dependent on eIF2alpha phosphorylation rather than the

    mTORC1/4E-BP1 pathway. PLoS One. 2010;5:e13910.

    29. Arnaudeau C, Lundin C, Helleday T. DNA double-strand breaks associated with replication

    forks are predominantly repaired by homologous recombination involving an exchange mechanism in

    mammalian cells. J Mol Biol. 2001;307:1235-45.

    30. Haaf T, Golub EI, Reddy G, Radding CM, Ward DC. Nuclear foci of mammalian Rad51

    recombination protein in somatic cells after DNA damage and its localization in synaptonemal

    complexes. Proc Natl Acad Sci U S A. 1995;92:2298-302.

    31. Tomicic MT, Kaina B. Topoisomerase degradation, DSB repair, p53 and IAPs in cancer cell

    resistance to camptothecin-like topoisomerase I inhibitors. Biochim Biophys Acta. 2013;1835:11-27.

    32. Pencreach E, Guerin E, Nicolet C, Lelong-Rebel I, Voegeli AC, Oudet P, et al. Marked activity of

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 31

    irinotecan and rapamycin combination toward colon cancer cells in vivo and in vitro is mediated

    through cooperative modulation of the mammalian target of rapamycin/hypoxia-inducible factor-

    1alpha axis. Clin Cancer Res. 2009;15:1297-307.

    33. Yin MB, Li ZR, Toth K, Cao S, Durrani FA, Hapke G, et al. Potentiation of irinotecan sensitivity

    by Se-methylselenocysteine in an in vivo tumor model is associated with downregulation of

    cyclooxygenase-2, inducible nitric oxide synthase, and hypoxia-inducible factor 1alpha expression,

    resulting in reduced angiogenesis. Oncogene. 2006;25:2509-19.

    34. Murono K, Tsuno NH, Kawai K, Sasaki K, Hongo K, Kaneko M, et al. SN-38 overcomes

    chemoresistance of colorectal cancer cells induced by hypoxia, through HIF1alpha. Anticancer Res.

    2012;32:865-72.

    35. Huang M, Gao H, Chen Y, Zhu H, Cai Y, Zhang X, et al. Chimmitecan, a novel 9-substituted

    camptothecin, with improved anticancer pharmacologic profiles in vitro and in vivo. Clin Cancer Res.

    2007;13:1298-307.

    36. Kang MJ, Jung SM, Kim MJ, Bae JH, Kim HB, Kim JY, et al. DNA-dependent protein kinase is

    involved in heat shock protein-mediated accumulation of hypoxia-inducible factor-1alpha in hypoxic

    preconditioned HepG2 cells. FEBS J. 2008;275:5969-81.

    37. Huang LE. Carrot and stick: HIF-alpha engages c-Myc in hypoxic adaptation. Cell Death Differ.

    2008;15:672-7.

    38. Liu Y, Bernauer AM, Yingling CM, Belinsky SA. HIF1alpha regulated expression of XPA

    contributes to cisplatin resistance in lung cancer. Carcinogenesis. 2012;33:1187-92.

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 32

    Figure legends

    Figure 1. Depletion of HIF-1α could reverse hypoxia-induced SN-38 resistance. (A)

    HIF-1α knockdown was achieved by transfection with a HIF-1α-targeting shRNA

    construct. Bel-7402 cells were transfected with either the empty-vector or the GFP-

    tagged shRNA-encoding constructs, followed by puromycin selection. Western blot

    analysis with specific antibodies to HIF-1α and visualization of GFP-positivity with

    an Olympus fluorview 1000 confocal microscope were used to select clones with

    stable knockdown of HIF-1α. (B) Bel-7402-HIF-1α-positive or Bel-7402-HIF-1α-

    negative cells obtained from (A) were treated with different concentrations of SN-38

    (0-1 μM) for 24 h under normoxic or hypoxic conditions, followed by subsequent

    proliferation inhibition assessment using the SRB assay. (C) Bel-7402 cells were

    pretreated with CoCl2 (200 μM) for 30 min to allow for functional inhibition of HIF-

    1α degradation under normoxic conditions. Cells were then exposed to different

    concentrations of SN-38 (0-1 μM) for 24 h under normoxic or hypoxic conditions,

    after which cell growth was determined by SRB assay. HIF-1α protein levels were

    also determined by western blot analysis. β-Actin was measured as the loading

    control. Each condition was studied in triplicate, and the error bars represent the

    standard deviation around the mean.

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 33

    Figure 2. TPZ enhanced the cytotoxicity elicited by topoisomerase I inhibitors under

    hypoxic conditions. (A) Bel-7402 cells were exposed to hypoxia or normoxia and

    treated with TPZ (0–10 μM) for 12 h. Protein levels of HIF-1α were detected by

    western blot analysis. β-Actin was used as the loading control. (B) Bel-7402 cells

    were exposed to hypoxia or normoxia and treated with TPZ (0–10 μM) for 6 h. An

    HRE-dependent reporter assay was used to determine the effect of TPZ on HIF-1α

    transcriptional activity. Five independent experiments were performed, and the values

    are expressed as the mean ± SD. *P < 0.05, **P < 0.01 and ***P < 0.001, compared

    to untreated controls in hypoxia. (C-F) Bel-7402 cells exposed to serial concentrations

    of TPZ, topoisomerase I inhibitors (SN-38/ TPT/ HCPT/ MONCPT), or the constant-

    ratio mixture of these two for 24 h under hypoxic condition. Cell growth was

    determined by the SRB assay. Each condition was studied in triplicate, and the error

    bars represent the standard deviation around the mean.

    Figure 3. TPZ enhanced the cytotoxicity elicited by SN-38 in five HCC cell lines

    under hypoxia. (A-E) Five HCC cell lines, Bel-7402, HepG2, SMMC-7721, Hep3B

    and Huh 7, were treated with TPZ (0–20 μM) and/or SN-38 (0–0.2 μM) for 24 h

    under hypoxia. Cell growth was determined by the SRB assay. The data are

    representative of three independent experiments and are shown as the mean ± SD. (F)

    Co-treatment with TPZ did not increase the toxicity of SN-38 in a normal human liver

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 34

    cell line. A normal liver cell line, Chang liver cells, were treated with TPZ (0–20 μM)

    and/or SN-38 (0–0.2 μM) for 24 h under hypoxia. Cell viability was determined by

    the SRB assay. Three independent experiments were performed, and the values are

    expressed as the mean ± SD.

    Figure 4. TPZ enhanced the apoptosis induced by SN-38 in Bel-7402. (A-D) Bel-

    7402 cells were exposed to SN-38 (0.1 μM) and/or TPZ (10 μM) for 12 h under

    hypoxic conditions. (A) Apoptotic bodies were visualized by confocal microscopy

    using DAPI staining. Scale bar, 50 μm. (B) Percentage of apoptosis was measured by

    flow cytometry using PI staining. Three independent experiments were performed,

    and the values are expressed as the mean ± SD. **P < 0.01, compared to group treated

    with TPZ or SN-38 alone. (C) Loss of mitochondrial membrane potential (ΔΨm) was

    measured by flow cytometry using the unique lipophilic cationic dye (JC-1). Three

    independent experiments were performed, and the values are expressed as the mean ±

    SD. *P < 0.05, compared to group treated with TPZ or SN-38 alone. (D) Percentage

    of apoptosis was measured by flow cytometry using the Annexin V/PI apoptosis

    detection kit. Three independent experiments were performed, and the values are

    expressed as the mean ± SD. **P < 0.01, compared to group treated with TPZ or SN-

    38 alone. (E) Bel-7402 cells were exposed to SN-38 (0.1μM) and/or TPZ (10 μM) for

    0-24 h under hypoxic conditions. Protein levels of PARP, Caspase 3, and cleaved-

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 35

    Caspase 3 were determined by western blot analysis. β-Actin was used as the loading

    control.

    Figure 5. The role of HIF-1α in the cellular apoptosis induced synergistically by SN-

    38 and TPZ. (A-D) Bel-7402 cells were exposed to SN-38 (0.1 μM) and/or TPZ (10

    μM) for 4 h under hypoxic conditions. (A) Protein levels of HIF-1α, HIF-1β, HIF-2α,

    p-ATM, ATM, p-Chk1, Chk1, p-Chk2, Chk2, Rad51 and γ-H2AX were determined by

    western blot analysis. β-Actin was used as the loading control. (B) Protein levels of

    HIF-1α and γ-H2AX were analyzed by confocal microscopy. Scale bar, 20 μm. (C)

    Total RNA was extracted and HIF-1α and VEGF mRNA expression was analyzed by

    RT-PCR, using GAPDH as a housekeeping gene. Untreated controls in hypoxia were

    scaled to 100%. Five independent experiments were performed, and the values are

    expressed as the mean ± SD. *P < 0.05, compared to group treated with TPZ or SN-38

    alone. (D) An HRE-dependent reporter gene assay was used to determine the effect of

    SN-38 and/or TPZ on HIF-1α transcriptional activity. Five independent experiments

    were performed, and the values are expressed as the mean ± SD. *P < 0.05, **P <

    0.01 and ***P < 0.001, compared to group treated with TPZ or SN-38 alone. (E) Bel-

    7402 cells were pre-incubated with CHX for 3 h in normoxic conditions and then

    placed in fresh medium and treated with or without 10 µM TPZ plus 100 nM SN-38

    for the indicated times under hypoxic conditions. The cells were harvested, and

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • 36

    lysates were immunoblotted with an anti-HIF-1α antibody. β-Actin was used as the

    loading control. (F) HIF-1α deficient and wild type Bel-7402 cells were both exposed

    to SN-38 (0.1 μM) for 2 h under hypoxic conditions and then placed in fresh medium

    for 6 h under hypoxic conditions. Protein levels of HIF-1α and γ-H2AX at different

    time points were analyzed by confocal microscopy. Scale bar, 20 μm.

    Figure 6. The synergistic effect of SN-38 and TPZ on Bel-7402 human xenograft

    models. (A-E) Nude mice bearing established Bel-7402 tumors were treated with

    CPT-11 at 2.5 mg/kg by daily i.p. injections, and/or TPZ at 25 mg/kg by i.p. injection

    every two days for 27 days. (A) Tumor volumes are expressed as the mean ± SEM.

    ***P < 0.001 versus vehicle-treated controls (n = 6 per group). (B) Relative tumor

    volumes are expressed as the mean ± SEM. **P < 0.01 versus vehicle-treated controls

    (n = 6 per group). (C) Body weights are expressed as the mean ± SEM. (D)

    Expression of PARP, Caspase 3, cleaved-Caspase 3 and HIF-1α extracted from tumor

    tissues from drug-administrated mice of the four groups were detected by western blot.

    β-Actin was used as the loading control. (E, F) Effects of CPT-11 and/or TPZ on

    expression levels of HIF-1α in Bel-7402 cell-derived tumors were determined by

    hematoxylin and eosin (H&E) staining and immunohistochemical and

    immunofluorescence analysis.

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/

  • Published OnlineFirst December 20, 2013.Mol Cancer Ther Tian-Yu Cai, Xiao-Wen Liu, Hong Zhu, et al.

    αHypoxia-Inducible Factor-1Topoisomerase I Inhibitors via Cooperative Modulation of Tirapazamine Sensitizes Hepatocellular Carcinoma Cells to

    Updated version

    10.1158/1535-7163.MCT-13-0490doi:

    Access the most recent version of this article at:

    Material

    Supplementary

    http://mct.aacrjournals.org/content/suppl/2013/12/23/1535-7163.MCT-13-0490.DC1

    Access the most recent supplemental material at:

    Manuscript

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

    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://mct.aacrjournals.org/content/early/2013/12/20/1535-7163.MCT-13-0490To request permission to re-use all or part of this article, use this link

    on June 21, 2021. © 2013 American Association for Cancer Research. mct.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 December 20, 2013; DOI: 10.1158/1535-7163.MCT-13-0490

    http://mct.aacrjournals.org/lookup/doi/10.1158/1535-7163.MCT-13-0490http://mct.aacrjournals.org/content/suppl/2013/12/23/1535-7163.MCT-13-0490.DC1http://mct.aacrjournals.org/cgi/alertsmailto:[email protected]://mct.aacrjournals.org/content/early/2013/12/20/1535-7163.MCT-13-0490http://mct.aacrjournals.org/

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