Butein inhibits NF-κB, AP-1 and Akt activation in adult T-cell leukemia/lymphoma

  • Authors:
    • Chie Ishikawa
    • Masachika Senba
    • Naoki Mori
  • View Affiliations

  • Published online on: June 2, 2017     https://doi.org/10.3892/ijo.2017.4026
  • Pages: 633-643
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Human T-cell leukemia virus type 1 (HTLV-1) is the causative agent of adult T-cell leukemia/lymphoma (ATLL) but there is no effective treatment for HTLV-1-associated diseases. Herein, we determined the effect of butein, a bioactive plant polyphenol, on cell growth, apoptosis and signaling pathways in HTLV-1-infected T-cell lines and on tumor growth in SCID mice. Treatment with butein caused a decrease in viability of HTLV-1-infected T-cell lines. T cells cultured with butein showed obvious apoptosis morphology, and cleavage of poly(ADP-ribose) polymerase with activation of caspase-3, -8 and -9. Pretreatment of cells with caspase inhibitor partially blocked butein-induced inhibition of cell viability. Butein also resulted in cell cycle arrest at G1 phase. Butein markedly downregulated the protein expression levels of CDK4, CDK6, cyclin D1, cyclin D2, cyclin E, survivin, XIAP, c-IAP2 and phospho-pRb. Butein also inhibited i) total and phospho-protein levels of IκB kinase (IKK)α and IKKβ, ii) degradation and phosphorylation of IκBα, iii) JunB and JunD, iv) total and phospho-protein levels of Akt, v) phosphorylation of RelA, vi) heat shock protein 90, and vii) DNA-binding activity of NF-κB and AP-1. In mice harboring ATLL xenograft tumors, butein caused a significant inhibition of tumor growth and reduced serum levels of soluble interleukin-2 receptor α chain and soluble cluster of differentiation 30. Considered together, the results indicated that butein has antiproliferative and proapoptotic properties through the suppression of NF-κB, AP-1 and Akt signaling in HTLV-1-infected T cells, both in vitro and in vivo, suggesting its therapeutic potential against HTLV-1-associated diseases including ATLL.

Introduction

The retrovirus, human T-cell leukemia virus type 1 (HTLV-1), is associated with adult T-cell leukemia/lymphoma (ATLL) and chronic inflammatory disorders such as HTLV-1-associated myelopathy/tropical spastic paraparesis (HAM/TSP), uveitis and arthritis (1). Globally, approximately 20 million people are estimated to be infected by HTLV-1, and 90% of them remain asymptomatic carriers during their lives (2). ATLL is a highly aggressive malignancy of mature CD4+ T cells and develops only after a long period of latency, usually several decades (1). Despite advances in the development of novel therapeutic agents, the prognosis of ATLL remains poor, and there is no definite cure for HAM/TSP (1). Many patients acquire resistance to chemotherapeutic agents, leading to treatment failures. A higher viral load in individuals with HTLV-1 infection increases their risk of developing ATLL and HAM/TSP (3,4). Therefore, reduction of the number of HTLV-1-infected T cells is crucial in the prevention and treatment of HTLV-1-associated diseases.

HTLV-1 induces oncogenesis through deregulation of selected cellular signaling pathways. In the virally infected T cells, activation of nuclear factor-κB (NF-κB), activator protein-1 (AP-1) and Akt results in upregulation of expression of a large number of cellular genes involved in cell proliferation, survival and metastasis (5,6). Thus, any new therapeutic agent must be designed to target multiple signal transduction pathways for maximum therapeutic benefits.

Natural phytochemicals, present in the human diet with cancer preventive and anticancer properties, are suitable alternatives (7). The plant polyphenol, butein (3,4,2′,4′-tetrahydroxychalone), has been traditionally used as an oriental medicine in Eastern Asia, and has various biological functions (8,9). Butein affects apoptosis, cell proliferation and the cell cycle in diverse cancers (8,9). The preliminary results of a clinical trial on the effects of butein-containing flavonoids in gastric cancer patients showed it was safe with good tolerability, and produced marked decrease in tumor size (10). Thus, butein seems to be a potent anticancer agent. Several previous reports have suggested the involvement of NF-κB, Akt, extracellular signal-regulated kinase, p38 kinase and p53 in the induction of cell cycle arrest at G1 or G2/M and apoptosis, and inhibition of metastasis by butein in various cell lines (8,9,1113). Therefore, we hypothesized that butein may act as a chemopreventive and/or chemotherapeutic agent against HTLV-1-associated diseases.

We report herein that butein inhibits cell growth and stimulates apoptosis of HTLV-1-infected T cells by inhibiting NF-κB, AP-1 and Akt activities and signaling pathways. Importantly, in a xenograft model, butein inhibited HTLV-1-infected T cell-induced tumor growth. Our study revealed for the first time, butein effect on HTLV-1-infected T cells both in vitro and in vivo, and the molecular mechanism by which butein exhibited anti-ATLL effects.

Materials and methods

Reagents used

Butein (cat. no. B3803) was purchased from Tokyo Chemical Industry Co. (Tokyo, Japan), dissolved in dimethyl sulfoxide (DMSO) and used for the treatment of cells. The broad spectrum caspase inhibitor, Z-VAD-FMK (cat. no. G7232), was purchased from Promega Corp. (Madison, WI, USA). Antibodies against survivin (cat. no. 2808), Akt (cat. no. 9272), IκB kinase (IKK)α (cat. no. 2682), IKKβ (cat. no. 2684), phospho-Akt (Thr308) (cat. no. 13038), phospho-Akt (Ser473) (cat. no. 4060), phospho-IκBα (Ser32 and 36) (cat. no. 9246), phospho-IKKα/β (Ser176/180 and Ser177/181) (cat. no. 2694), RelA (cat. no. 8242), phospho-RelA (Ser536) (cat. no. 3033), cleaved caspase-3 (cat. no. 9664), -8 (cat. no. 9496), -9 (cat. no. 9501) and poly(ADP-ribose) polymerase (PARP) (cat. no. 9541) were purchased from Cell Signaling Technology, Inc. (Beverly, MA, USA). Antibodies against Bcl-2 (cat. no. MS-597), CDK2 (cat. no. MS-617), CDK4 (cat. no. MS-299), CDK6 (cat. no. MS-398), cyclin E (cat. no. MS-870), retinoblastoma protein (pRb) (cat. no. MS-107) and actin (cat. no. MS-1295) were obtained from Neomarkers, Inc. (Fremont, CA, USA). Antibodies against XIAP (cat. no. M044-3), cyclin D1 (cat. no. K0062-3), heat shock protein (HSP)70 (cat. no. SR-812C), phospho-pRb (Ser780) (cat. no. M054-3S) were purchased from Medical & Biological Laboratories, Co. (Aichi, Japan). Antibodies against cyclin D2 (cat. no. sc-593), c-IAP2 (cat. no. sc-7944), IκBα (cat. no. sc-371), JunB (cat. no. sc-46) and JunD (cat. no. sc-74), and NF-κB subunits p50 (cat. no. sc-114X), p52 (cat. no. sc-298X), RelA (cat. no. sc-109X), c-Rel (cat. no. sc-70X) and RelB (cat. no. sc-226X), and AP-1 subunits c-Fos (cat. no. sc-52X), FosB (cat. no. sc-48X), Fra-1 (cat. no. sc-605X), Fra-2 (cat. no. sc-604X), c-Jun (cat. no. sc-45X), JunB (cat. no. sc-46X) and JunD (cat. no. sc-74X) for supershift assay were from Santa Cruz Biotechnology, Inc. (Santa Cruz, CA, USA). Antibody against HSP90 (cat. no. 610418) was from BD BioSciences (San Jose, CA, USA).

Cell culture

The HTLV-1-infected MT-4 and HUT-102, and ATLL-derived TL-OmI T-cell lines, were cultured in RPMI-1640 medium (cat. no. 30264-56, Nacalai Tesque, Inc., Kyoto, Japan) supplemented with 10% heat-inactivated fetal bovine serum (Biological Industries, Kibbutz Beit Haemek, Israel) and 1% penicillin/streptomycin (cat. no. 09367-34, Nacalai Tesque, Inc.). Cells were maintained in a 37°C, 5% CO2 humidified incubator. The study also included peripheral blood mononuclear cells (PBMCs) obtained from a healthy donor. PBMC proliferation was induced with 20 µg/ml of phytohemagglutinin (PHA) (cat. no. L8754, Sigma-Aldrich Co., St. Louis, MO, USA).

Cell proliferation and cytotoxic assay

The cell proliferative and toxic effects of butein were determined by the water-soluble tetrazolium (WST)-8 uptake method. Cells were plated (1×104 cells per well) in 96-well microtiter plates in triplicate and treated with various concentrations of butein. After incubation for 24–72 h, 10 µl of the WST-8 reagent (cat. no. 07553-44, Nacalai Tesque, Inc.) was added to each well. After 4 h, WST-8 reduction was measured at 450 nm using a 680 XR microplate reader (Bio-Rad Laboratories, Inc., Hercules, CA, USA). Values were normalized to untreated control samples. The 50% growth inhibitory concentration (IC50) value was calculated by dotting the data points to a logistic curve using the CalcuSym software (version 2.0; Biosoft, Cambridge, UK).

Analysis of cell apoptosis

Cells were treated with vehicle or butein for 24–72 h and then permeabilized by incubation on ice for 20 min with 100 µg/ml of digitonin, and treated with phycoerythrin-conjugated APO2.7 antibody (cat. no. IM2088, Beckman Coulter, Inc., Marseille, France) for 15 min at room temperature. After staining with APO2.7 antibody, apoptosis was determined by Epics XL flow cytometry (Beckman Coulter, Inc., Brea, CA, USA). In addition, for analysis of morphological changes in nuclei, cells were stained by 10 µg/ml of Hoechst 33342 (cat. no. 346-07951, Dojindo Molecular Technologies, Inc., Kumamoto, Japan) and observed under a Leica DMI6000 microscope (Leica Microsystems, Wetzlar, Germany).

In vitro measurement of caspase Activity

Caspase activity was measured using Colorimetric Caspase Assay kits (cat. nos. 4800, 4805 and 4810, Medical & Biological Laboratories, Co.). Briefly, cell extracts were recovered using the cell lysis buffer supplied with the kit and assessed for caspase-3, -8 and -9 activities using colorimetric probes. The assay kits are based on detection of chromophore ρ-nitroanilide after cleavage from caspase-specific labeled substrates. Colorimetric readings were performed in an automated microplate reader.

Cell cycle analysis

Cells were stained with the CycleTEST Plus DNA Reagent kit (cat. no. 340242, Becton-Dickinson Immunocytometry Systems, San Jose, CA, USA). The cell cycle distribution was analyzed for 10,000 collected cells by an Epics XL flow cytometry that uses the MultiCycle software (version 3.0; Phoenix Flow Systems, San Diego, CA, USA). The population of nuclei at each phase of the cell cycle was determined and apoptotic cells with hypodiploid DNA content were detected in the sub-G0/G1 region.

Western blot analysis

Whole cell extracts were prepared by subjecting butein-treated cells to lysis in lysis buffer [62.5 mM Tris-HCl (pH 6.8) (cat. no. 35434-21, Nacalai Tesque, Inc.), 2% sodium dodecyl sulfate (cat. no. 31607-65, Nacalai Tesque, Inc.), 10% glycerol (cat. no. 17045-65, Nacalai Tesque, Inc.), 6% 2-mercaptoethanol (cat. no. 21438-82, Nacalai Tesque, Inc.) and 0.01% bromophenol blue (cat. no. 021-02911, Wako Pure Chemical Industries, Osaka, Japan)]. Lysates were spun to remove insoluble material. The supernatants were collected and kept at -80°C. Protein concentrations were determined using a commercial kit (DC Protein Assay, cat. no. 5000116JA, Bio-Rad Laboratories, Inc.). Equal amounts of protein extracts (20 µg) were resolved on sodium dodecyl sulfate-polyacrylamide gels (SDS-PAGE). After electrophoresis, the proteins were electrotransferred onto polyvinylidene difluoride membranes (cat. no. IPVH00010EMD, Millipore, Darmstadt, Germany), blotted with the relevant antibodies. Finally, blots were hybridized with horseradish peroxidase-conjugated secondary anti-mouse (cat. no. 7076, Cell Signaling Technology, Inc.) or anti-rabbit IgG antibody (cat. no. 7074, Cell Signaling Technology, Inc.) and developed using an enhanced chemiluminescence reagent (cat. no. RPN2232, Amersham Biosciences Corp., Piscataway, NJ, USA).

RT-PCR

TRIzol reagent (cat. no. 15596026, Invitrogen Life Technologies, Carlsbad, CA, USA) was used to extract total RNA from cells. RNA (1 µg) was reverse transcribed into cDNA using a PrimeScript RT-PCR kit (cat. no. RR014A, Takara Bio Inc., Otsu, Japan). The sequence-specific primers used for RT-PCR were as follows: for IκBα, forward 5′-GCCTGGACTCCATGAAAGAC-3′, reverse 5′-CAAGTG GAGTGGAGTCTGCTGCAGGTTGTT-3′; and for GAPDH, forward 5′-GCCAAGGTCATCCATGACAACTTTGG-3′, reverse 5′-GCCTGCTTCACCACCTTCTTGATGTC-3′. PCR amplifications were performed in a programmable thermal cycler. Results were analyzed at 30 cycles to obtain results at exponential levels of amplification.

Electrophoretic mobility shift assay (EMSA)

To determine NF-κB and AP-1 activation, we prepared nuclear extracts from butein-treated cells and performed EMSA, as previously described (14). Nuclear extracts were incubated with 32P-labeled probes. The top strand sequences of the oligonucleotide probes or competitors were as follows: for a typical NF-κB element of the interleukin-2 receptor α chain (IL-2Rα) gene, 5′-GATCCGGCAGGGGAATCTCCCTCTC-3′ and for the consensus AP-1 element of the IL-8 gene, 5′-GATCGTGATGACTCAGGTT-3′. The above underlined sequences are the NF-κB and AP-1 binding sites, respectively. In competition experiments, nuclear extracts were preincubated for 15 min with 100-fold excess of unlabeled oligonucleotides. For supershift assays, nuclear extracts were incubated with antibodies against NF-κB or AP-1 subunits for 45 min at room temperature before the complex was analyzed by EMSA. The dried gels were visualized.

Xenograft tumor model

Five-week-old female C.B-17/Icr-severe combined immune deficient (SCID) mice were obtained from Kyudo, Co. (Tosu, Japan). To induce malignancy, 1×107 HUT-102 cells suspended in sterile RPMI-1640 medium (200 µl) were inoculated subcutaneously into the postauricular region of SCID mice. The mice were divided at random into two treatment groups (n=6, each). Butein was solubilized in soybean oil (cat. no. 190-03776, Wako Pure Chemical Industries) and administered at a dose of 0.7 mg/kg intraperitoneally three times a week, and the treatment was continued for 27 days, beginning on the day after cell inoculation. The control group received the vehicle only. The tumor diameter was measured weekly with a shifting caliper and tumor volume was calculated. The mice were also weighed weekly, beginning on day 0. All mice were sacrificed on day 28. The tumors were excised and their weight measured. At the end of the study, blood samples were collected and the sera were separated by centrifuging blood and then stored at −80°C until assayed for secreted soluble IL-2Rα [soluble cluster of differentiation 25 (sCD25)] and sCD30. This experiment was performed according to the Guidelines for Animal Experimentation of the University of the Ryukyus (Nishihara, Japan), and was approved by the Animal Care and Use Committee of the University of the Ryukyus (reference no. 5837).

Morphological analysis of tumor tissues and terminal deoxynucleotidyl transferase deoxyuridine triphosphate nick end labeling (TUNEL) assay

The tumor specimens were collected from the control mice group and 0.7 mg/kg butein group, fixed in formalin (Wako Pure Chemical Industries) solution, dehydrated through graded ethanol series (Japan Alcohol Selling Co., Tokyo, Japan) and embedded in paraffin (cat. no. 09620, Sakura Finetek Japan Co., Tokyo, Japan). The paraffin-embedded specimens of ATLL tumors were stained with hematoxylin and eosin (H&E, cat. nos. 234-12 and 1159350025, Merck, Darmstadt, Germany) and examined histologically. Analysis of DNA fragmentation by TUNEL testing was performed using a commercial kit (cat. no. 11684817910, Roche Applied Science, Penzberg, Germany). Cells were examined under a light microscope (Axioskop 2 Plus) with an Achroplan 40×/0.65 lens (both from Zeiss, Hallbergmoos, Germany). Images were acquired with an AxioCam 503 color and AxioVision LE64 software (Zeiss).

Biomarker analysis

Serum concentrations of human sCD25 (cat. no. DR2A00, R&D Systems, Inc., Minneapolis, MN, USA) and human sCD30 (cat. no. SK00582-01, Aviscera Bioscience, Inc., Santa Clara, CA, USA) were measured by enzyme-linked immunosorbent assay (ELISA), according to the protocol supplied by the manufacturer.

Statistical analysis

Values are shown as mean ± standard deviation (SD). Data of two groups were compared by the Student's t-test. Differences were considered significant at P<0.05.

Results

Effects of butein on cell proliferation and cytotoxicity of HTLV-1-infected T-cell lines

To determine the in vitro effects of butein in ATLL cells, two HTLV-1-infected T-cell lines (MT-4 and HUT-102) and an ATLL-derived T-cell line (TL-OmI) were cultured in the presence of butein (1.56–100 µM). As shown in Fig. 1, 24–72 h culture with butein significantly reduced cell proliferation and viability in a concentration- and time-dependent manner, with IC50 values at 72 h of 7.0 µM for HUT-102, 7.9 µM for MT-4 and 9.8 µM for TL-OmI. In contrast, the IC50 was 27.8 µM for normal PBMCs. Butein suppressed the proliferative response of PBMCs to the T cell mitogen PHA, with IC50 of 18.5 µM. Although the IC50 values for HTLV-1-infected T-cell lines were lower than those for normal PBMCs, obvious selective cytotoxic activities of butein on HTLV-1-infected T-cell lines have not been demonstrated.

Butein simultaneously induces cell apoptosis and cell cycle arrest

First, the effect of butein on cell morphology was examined using microscopy. Cells cultured with butein for 48 h demonstrated morphologies characteristic of apoptosis, such as chromatin condensation and nuclear fragmentation (Fig. 2A). To further investigate the mode of butein-induced cell death, HTLV-1-infected T-cell lines were treated with 6.3–25 µM butein for 24 h. Butein induced concentration-dependent increases in the sub-G0/G1 population (Fig. 3A and B). We also assessed the expression of 38-kDa mitochondrial membrane antigen APO2.7 during apoptosis (15). Apoptotic cells (APO2.7-positive) increased after treatment with butein concentration- and time-dependently (Fig. 2B and C). Collectively, these results indicate that butein induces apoptosis of HTLV-1-infected T-cell lines. In addition to apoptosis, the same experiment also showed that butein dose-dependently increased the percentage of cells in the G0/G1 phase and also reduced the percentage of cells in the S phase (Fig. 3A and B). Considered together, the results showed the arrest of more cells at the G0/G1 phase of the cell cycle.

Butein induces caspase activation

The apoptotic process is executed by a member of the highly conserved cysteine proteases, caspases. Initiator caspase-8 and -9 or effector caspase-3 are synthesized as inactive proenzymes and become finally activated by proteolytic cleavage (16). To identify the mechanism underlying butein-induced apoptosis of HTLV-1-infected T cells, we investigated cleavage of caspase-3, -8 and -9, and also PARP, a substrate for caspase-3. Exposure of MT-4 and HUT-102 cells to butein (12.5–100 µM for 48 h) caused concentration-dependent increase in the cleavage of caspase-3, -8 and -9, and PARP (Fig. 4A), as well as increase in activities of caspase-3, -8 and -9 (Fig. 4B). To further explore the significance of the latter activation, a broad spectrum caspase inhibitor, Z-VAD-FMK, was used to suppress the effect of butein. Pretreatment of cells with Z-VAD-FMK partially attenuated butein-induced inhibition of cell viability (Fig. 4C), suggesting that butein-mediated cytotoxic effect is mediated, at least in part, through a caspase-dependent apoptosis pathway.

Effects of butein on regulatory proteins involved in cell cycle regulation and apoptosis

Cell cycle progression is tightly regulated by interaction between CDKs and cyclins (17). To identify the molecular mechanism underlying butein-induced G0/G1 arrest, cells were cultured with butein (12.5–100 µM) for 48 h. They were harvested for protein extraction and western blot analysis. As shown in Fig. 5A, levels of CDK4, CDK6, cyclin D2 and cyclin E proteins were significantly decreased in butein-treated MT-4 and HUT-102 cells, compared to DMSO-treated cells. Butein also inhibited the expression of cyclin D1 in HUT-102 but not in MT-4 cells. However, butein had no significant effect on CDK2 protein level. Furthermore, butein induced dephosphorylation of pRb protein (Fig. 5A, top panels) and changed the hyperphosphorylated form to a hypophosphorylated form of pRB (Fig. 5A, second panels). The deregulation of apoptosis is often studied by the expression of anti-apoptotic proteins. There was a dose-dependent decrease in the protein expression of survivin and XIAP, but not Bcl-2, in butein-treated MT-4 and HUT-102 cells. Butein also inhibited the expression of c-IAP2 in HUT-102 but not in MT-4 cells (Fig. 5B). The effects of butein seem to slightly differ dependent on cell lines.

Butein inhibits NF-κB in HTLV-1-infected T-cell lines

NF-κB forms a family of transcription factors that regulate numerous biological processes, including immune response, inflammation, cell growth, survival and development. The NF-κB proteins are usually sequestered in the cytoplasm by a family of inhibitors, including IκBα. IκBα degradation is mediated through its phosphorylation by IKK, a trimeric complex composed of two catalytic subunits, IKKα and IKKβ, and a regulatory subunit, IKKγ (18). Incubation of MT-4 and HUT-102 cells with butein resulted in significant dose-dependent inhibition of phosphorylation and upregulation of IκBα protein (Fig. 6A). To determine whether butein-mediated upregulation of IκBα is due to transcriptional activation of IκBα, we examined the expression of IκBα mRNA in cells treated with butein by RT-PCR. As shown in Fig. 6B, butein treatment had no influence on IκBα mRNA, suggesting that butein induces the inhibition of degradation of IκBα protein. Next, we measured total and phospho-protein levels of IKKα and IKKβ to evaluate the possible inhibitory mechanism of butein on IκBα protein phosphorylation. Immunoblot analysis demonstrated butein dose-dependent suppression of phospho-protein levels of IKKα and IKKβ in MT-4 and HUT-102 cells. Total protein levels of both IKKα and IKKβ were reduced in HUT-102 cells, while total protein level of IKKβ alone was reduced in MT-4 cells (Fig. 6A). Thus, the effects of butein seem to slightly differ dependent on cell lines.

NF-κB is a complex of proteins, in which various combinations of NF-κB proteins constitute active NF-κB heterodimers that bind to specific DNA sequences. Thus, to show that the band visualized by EMSA in HTLV-1-infected T-cell lines was indeed NF-κB, nuclear extracts from MT-4 and HUT-102 cells were incubated with antibodies to the NF-κB subunits and analyzed by EMSA. The results showed shifts of the bands to higher molecular masses (Fig. 6C), suggesting that the activated complex consisted of p50, RelA and RelB. The addition of excess unlabeled NF-κB caused complete disappearance of the band, whereas the addition of AP-1 oligonucleotide had no effect on the DNA binding. We also investigated whether butein could inhibit constitutive NF-κB in HTLV-1-infected T-cell lines. MT-4 and HUT-102 cells were cultured in the presence of different concentrations of butein for 24 h and then analyzed for NF-κB activation. Butein dose-dependently inhibited constitutive NF-κB activation in HTLV-1-infected T-cell lines (Fig. 6D). These results indicate that butein can suppress constitutively active NF-κB in HTLV-1-infected T cells.

Butein suppresses AP-1 activity

AP-1 is a group of dimeric transcription factor complexes composed of members of the Fos (c-Fos, FosB, Fra-1 and Fra-2) and Jun (c-Jun, JunB and JunD) families, which play central roles in the proliferation and transformation of T cells (6). EMSA showed supershifted band for AP-1, as detected with specific antibodies to JunB and JunD in MT-4 and HUT-102 cells (Fig. 7A). As shown in Fig. 7B, butein produced a significant and dose-dependent suppression of AP-1 activation in HTLV-1-infected T-cell lines. As shown in Fig. 7A, JunB and JunD are functional components of AP-1 transcription factor in HTLV-1-infected T-cell lines. The experiment demonstrated that butein inhibited AP-1 signaling pathway through dose-dependent suppression of JunB and JunD protein levels (Fig. 7C).

Butein inhibits Akt and HSP90 protein expression in HTLV-1-infected T-cell lines

Akt is a component of an essential pathway for cell survival and growth during development and carcinogenesis. It regulates cell survival by directly targeting XIAP and by indirectly modulating survivin levels (19,20). Akt also regulates cell cycle and proliferation by indirectly modulating the levels of cyclin D1, cyclin D2 and cyclin E (21). Since butein decreased the levels of XIAP, survivin, cyclin D1, cyclin D2 and cyclin E in the treated cells, we determined its effect on Akt protein. Butein at 50–100 µM for MT-4 and 12.5–100 µM for HUT-102 cells, caused a dose-dependent inhibition of Akt protein expression. Treatment of MT-4 cells with butein at 50–100 µM and HUT-102 cells with butein at 12.5–100 µM caused a dose-dependent inhibition in Akt phosphorylation at Thr308. Butein at 12.5–100 µM for MT-4 and 25–100 µM for HUT-102 cells, also inhibited the phosphorylation of Akt at Ser473 (Fig. 8). RelA is phosphorylated at Ser536 by Akt and such phosphorylation enhances RelA trans-activation potential (22). Butein also caused inhibition of RelA phosphorylation (Fig. 8). The molecular chaperone HSP90 is involved in the stabilization and conformational maturation of many signaling proteins that are deregulated in cancers. HSP90 client proteins include signaling proteins (IKKs and Akt) and cell cycle regulators (CDKs) (23). Therefore, we investigated the effect of butein on the expression of HSP90. As shown in Fig. 8, butein reduced HSP90 protein levels, but increased HSP70 levels.

Butein suppresses growth of HTLV-1-infected T-cell line in SCID mice

Butein significantly inhibited tumor growth in SCID mice inoculated with HTLV-1-infected HUT-102 cells (Fig. 9A, left panels). In the control group, the average tumor volume of 1,545 mm3 was reached in 28 days after HUT-102 cell inoculation (Fig. 9B, left panel). At the same time point, the average tumor volume in the butein-treated group was 727 mm3. Tumor tissues were removed, photographed (Fig. 9A, right panel), and weighed. The mean tumor weight of butein-treated mice was significantly lower than that of vehicle-treated mice (Fig. 9B, middle panel). In addition, the body weight of mice remained stable with no significant differences between the two groups (Fig. 9B, right panel). At post-inoculation day 28, quantitative ELISA used to determine the circulating levels of surrogate tumor markers sCD25 (24) and sCD30 (25) secreted by HUT-102 tumor xenografts showed 42% and 59% decrease in their levels, respectively, in butein-treated mice, compared with the control group (Fig. 9C). Considered together, our results showed that treatment with butein resulted in significant decreases in serum sCD25 and sCD30 of SCID mice.

Finally, H&E and TUNEL staining assays were performed to assess whether the inhibitory effect on tumor growth induced by butein was due to an increase in apoptotic cell death. H&E staining showed a significant increase in apoptotic tumor cells in butein-treated mice, compared with the control group, which were characterized by cytoplasmic condensation, chromatin hyperchromatism and condensation, and nuclear fragmentation. TUNEL assays confirmed these findings and demonstrated increased number of TUNEL-positive tumor cells in butein-treated mice (Fig. 9D).

Discussion

ATLL is classified into four clinical subtypes based on the clinical manifestations: acute, lymphoma, chronic and smoldering (26). Acute, lymphoma and unfavorable chronic types are considered the aggressive forms of ATLL, while favorable chronic and smoldering types are indolent ATLL (26). Various dietary and lifestyle factors may be involved in the development and progression of ATLL. Aggressive ATLL carries poor prognosis, with only 24% for 3-year overall survival rate in patients on standard combination chemotherapy (27). Therefore, there is a great need for novel therapeutic approaches for ATLL. Chemotherapeutic agents are usually highly toxic to normal tissues. To reduce the number of HTLV-1-infected T cells, natural products are suitable alternatives since they have minimal side-effects compared to synthetic drugs.

In this study, we showed that butein decreased cell viability of HTLV-1-infected T-cell lines. The ability of the cells to divide is regulated by two classes of molecules, CDKs, a family of serine/threonine kinases and their binding partners, cyclins. The progress of the cells division process occurs through the interaction of various cyclins with their respective CDK subunits (28). Quiescent cells enter the cell cycle after mitogenic stimulation and upregulation of cyclins D and E during the G1 phase of the cell cycle (29). Cyclin D is associated with CDK4 and CDK6 (30), while cyclin E is associated with CDK2 (31). Thus, cyclins and CDKs play important roles in DNA synthesis and cell division. In contrast, pRb controls cell cycle progression at the G1 to S transition in response to certain signals for growth inhibition (32). pRb is phosphorylated by catalytic subunits of CDK4/CDK6 and CDK2 complexed with specific regulatory subunits of cyclins D and E, respectively. Treatment of HTLV-1-infected T-cell lines with butein reduced the expression levels of cyclin D1, D2, E, CDK4, CDK6 and phospho-pRb. Based on these findings, it is highly possible that butein causes G1 arrest by blocking cyclin D-CDK4/6- and cyclin E-CDK2-mediated pRb phosphorylation, which is crucial to the progression of the cell cycle at the G1 to S transition.

The process of apoptosis is controlled by various interrelated pathways, which ensure that caspases, the proteolytic initiators, and executioners of apoptosis, are triggered only in cells requiring termination. We have shown in the present study the activation of caspase-3, -8 and -9 during the application of butein and that the addition of a general caspase inhibitor partially decreased butein-mediated cell death. These data provide evidence that butein-induced apoptosis was at least in part caspase-dependent in HTLV-1-infected T-cell lines. Butein also downregulated the protein expression of survivin, XIAP and c-IAP2, which are possible upstream factors of caspase proteins.

NF-κB can modulate the transcriptional activation of genes associated with cell proliferation, metastasis, tumor promotion and suppression of apoptosis (33,34). Hence, agents capable of suppressing NF-κB activation are therapeutically promising in inhibiting carcinogenesis. Our results suggest that the effects of butein on NF-κB were mediated through inhibition of phosphorylation and the associated subsequent proteolysis of IκBα since butein blocked phosphorylation and degradation of IκBα. Furthermore, butein inhibited total and phospho-protein levels of IKKα and IKKβ, and phosphorylation of RelA. Treatment of HTLV-1-infected T-cell lines with butein further inhibited NF-κB-regulated gene products, survivin, XIAP, c-IAP2, cyclin D1, cyclin D2, CDK4 and CDK6.

Akt occupies a key regulatory node in the phosphoinositide 3-kinase pathway, below which the pathway branches significantly to influence a wide range of cellular processes that promote cell cycle progression, cell growth and resistance to apoptosis by modulating the levels of survivin, XIAP, cyclin D1, cyclin D2 and cyclin E (1921). Our results demonstrated that butein decreased total and phospho-protein levels of Akt in HTLV-1-infected T-cells. Akt is known to phosphorylate RelA at Ser536 through an IKKα- or IKKβ-dependent mechanism (35). Butein may inhibit phosphorylation of RelA through inactivation of Akt and IKKs. In addition, AP-1 regulates the expression and function of various cell cycle regulators, such as cyclin D1 and cyclin E (36). JunB is also controlled by NF-κB (37). Thus, NF-κB and AP-1 can also collaborate in ATLL, depending on a variety of factors. Butein inhibited AP-1 signaling pathway via reduction of JunB and JunD. Considered together, these findings suggest that the effects of butein are mediated through a diverse range of molecular targets, which collectively result in inhibition of cell growth and survival.

Of note, butein profoundly inhibited HSP90 protein. Previous studies showed that HSP90 inhibitors simultaneously inactivated and destabilized several oncoproteins (e.g., IKKs, Akt and CDKs) in cancer cells (23,38). On the other hand, inhibition of HSP90 alone induces upregulation of another chaperone, HSP70 (38). Our results also showed that butein increased the expression of HSP70. HSP70 is known to be anti-apoptotic, and therefore the induction of HSP70 may ultimately limit the efficacy of HSP90 inhibitors (38). Collectively, these results suggest that butein also inhibits HSP90, thereby affecting various signaling pathways involved in the regulation of cell growth and survival.

To establish the relevance of these in vitro findings to the intact animal, we implanted SCID mice with HTLV-1-infected T cell line HUT-102. Treatment with butein significantly decreased HUT-102 growth in these mice and significantly reduced serum levels of sCD25 and sCD30 without affecting body weight. The results also showed upregulation of TUNEL-positive cells undergoing early apoptosis in tissues of mice treated with butein, relative to the control mice.

In conclusion, we demonstrated in the present study the effects of butein, a plant polyphenol, on the growth of HTLV-1-infected T-cell lines in both in vitro and in an in vivo animal model. Although this study warrants further investigation on ATLL primary samples, the results suggest that butein is a potentially valuable chemopreventive and/or chemotherapeutic agent against ATLL.

Acknowledgments

We thank Fujisaki Cell Center, Hayashibara Biochemical Laboratories, Inc. (Okayama, Japan) for providing HUT-102. This work was supported in part by JSPS KAKENHI grant numbers, 25830085 and 25461428, and a grant from the Mishima Kaiun Memorial Foundation.

References

1 

Gonçalves DU, Proietti FA, Ribas JG, Araújo MG, Pinheiro SR, Guedes AC and Carneiro-Proietti AB: Epidemiology, treatment, and prevention of human T-cell leukemia virus type 1-associated diseases. Clin Microbiol Rev. 23:577–589. 2010. View Article : Google Scholar : PubMed/NCBI

2 

Gessain A and Cassar O: Epidemiological aspects and world distribution of HTLV-1 infection. Front Microbiol. 3:3882012. View Article : Google Scholar : PubMed/NCBI

3 

Okayama A, Stuver S, Matsuoka M, Ishizaki J, Tanaka G, Kubuki Y, Mueller N, Hsieh CC, Tachibana N and Tsubouchi H: Role of HTLV-1 proviral DNA load and clonality in the development of adult T-cell leukemia/lymphoma in asymptomatic carriers. Int J Cancer. 110:621–625. 2004. View Article : Google Scholar : PubMed/NCBI

4 

Nagai M, Usuku K, Matsumoto W, Kodama D, Takenouchi N, Moritoyo T, Hashiguchi S, Ichinose M, Bangham CR, Izumo S, et al: Analysis of HTLV-I proviral load in 202 HAM/TSP patients and 243 asymptomatic HTLV-I carriers: High proviral load strongly predisposes to HAM/TSP. J Neurovirol. 4:586–593. 1998. View Article : Google Scholar

5 

Mori N: Cell signaling modifiers for molecular targeted therapy in ATLL. Front Biosci (Landmark Ed). 14:1479–1489. 2009. View Article : Google Scholar

6 

Hall WW and Fujii M: Deregulation of cell-signaling pathways in HTLV-1 infection. Oncogene. 24:5965–5975. 2005. View Article : Google Scholar : PubMed/NCBI

7 

Surh Y-J: Cancer chemoprevention with dietary phytochemicals. Nat Rev Cancer. 3:768–780. 2003. View Article : Google Scholar : PubMed/NCBI

8 

Yadav VR, Prasad S, Sung B and Aggarwal BB: The role of chalcones in suppression of NF-κB-mediated inflammation and cancer. Int Immunopharmacol. 11:295–309. 2011. View Article : Google Scholar

9 

Gupta SC, Kim JH, Prasad S and Aggarwal BB: Regulation of survival, proliferation, invasion, angiogenesis, and metastasis of tumor cells through modulation of inflammatory pathways by nutraceuticals. Cancer Metastasis Rev. 29:405–434. 2010. View Article : Google Scholar : PubMed/NCBI

10 

Lee S-H, Choi W-C, Kim K-S, Park J-W, Lee S-H and Yoon S-W: Shrinkage of gastric cancer in an elderly patient who received Rhus verniciflua Stokes extract. J Altern Complement Med. 16:497–500. 2010. View Article : Google Scholar : PubMed/NCBI

11 

Woo S-M, Choi YK, Kim AJ, Cho S-G and Ko S-G: p53 causes butein-mediated apoptosis of chronic myeloid leukemia cells. Mol Med Rep. 13:1091–1096. 2016.

12 

Zhang L, Yang X, Li X, Li C, Zhao L, Zhou Y and Hou H: Butein sensitizes HeLa cells to cisplatin through the AKT and ERK/p38 MAPK pathways by targeting FoxO3a. Int J Mol Med. 36:957–966. 2015.PubMed/NCBI

13 

Bai X, Ma Y and Zhang G: Butein suppresses cervical cancer growth through the PI3K/AKT/mTOR pathway. Oncol Rep. 33:3085–3092. 2015.PubMed/NCBI

14 

Mori N and Prager D: Transactivation of the interleukin-1alpha promoter by human T-cell leukemia virus type I and type II Tax proteins. Blood. 87:3410–3417. 1996.PubMed/NCBI

15 

Zhang C, Ao Z, Seth A and Schlossman SF: A mitochondrial membrane protein defined by a novel monoclonal antibody is preferentially detected in apoptotic cells. J Immunol. 157:3980–3987. 1996.PubMed/NCBI

16 

Khan N, Afaq F and Mukhtar H: Apoptosis by dietary factors: The suicide solution for delaying cancer growth. Carcinogenesis. 28:233–239. 2007. View Article : Google Scholar

17 

Lim S and Kaldis P: Cdks, cyclins and CKIs: Roles beyond cell cycle regulation. Development. 140:3079–3093. 2013. View Article : Google Scholar : PubMed/NCBI

18 

Hayden MS and Ghosh S: Shared principles in NF-kappaB signaling. Cell. 132:344–362. 2008. View Article : Google Scholar : PubMed/NCBI

19 

Dan HC, Sun M, Kaneko S, Feldman RI, Nicosia SV, Wang H-G, Tsang BK and Cheng JQ: Akt phosphorylation and stabilization of X-linked inhibitor of apoptosis protein (XIAP). J Biol Chem. 279:5405–5412. 2004. View Article : Google Scholar

20 

Vaira V, Lee CW, Goel HL, Bosari S, Languino LR and Altieri DC: Regulation of survivin expression by IGF-1/mTOR signaling. Oncogene. 26:2678–2684. 2007. View Article : Google Scholar

21 

Zhang X, Tang N, Hadden TJ and Rishi AK: Akt, FoxO and regulation of apoptosis. Biochim Biophys Acta. 1813:1978–1986. 2011. View Article : Google Scholar : PubMed/NCBI

22 

Sizemore N, Leung S and Stark GR: Activation of phosphatidylinositol 3-kinase in response to interleukin-1 leads to phosphorylation and activation of the NF-kappaB p65/RelA subunit. Mol Cell Biol. 19:4798–4805. 1999. View Article : Google Scholar : PubMed/NCBI

23 

Kamal A, Boehm MF and Burrows FJ: Therapeutic and diagnostic implications of Hsp90 activation. Trends Mol Med. 10:283–290. 2004. View Article : Google Scholar : PubMed/NCBI

24 

Kamihira S, Atogami S, Sohda H, Momita S, Yamada Y and Tomonaga M: Significance of soluble interleukin-2 receptor levels for evaluation of the progression of adult T-cell leukemia. Cancer. 73:2753–2758. 1994. View Article : Google Scholar : PubMed/NCBI

25 

Nishioka C, Takemoto S, Kataoka S, Yamanaka S, Moriki T, Shoda M, Watanabe T and Taguchi H: Serum level of soluble CD30 correlates with the aggressiveness of adult T-cell leukemia/lymphoma. Cancer Sci. 96:810–815. 2005. View Article : Google Scholar : PubMed/NCBI

26 

Shimoyama M: Diagnostic criteria and classification of clinical subtypes of adult T-cell leukaemia-lymphoma. A report from the Lymphoma Study Group (1984–87). Br J Haematol. 79:428–437. 1991. View Article : Google Scholar : PubMed/NCBI

27 

Tsukasaki K, Utsunomiya A, Fukuda H, Shibata T, Fukushima T, Takatsuka Y, Ikeda S, Masuda M, Nagoshi H, Ueda R, et al Japan Clinical Oncology Group Study JCOG9801: VCAP-AMP-VECP compared with biweekly CHOP for adult T-cell leukemia-lymphoma: Japan Clinical Oncology Group Study JCOG9801. J Clin Oncol. 25:5458–5464. 2007. View Article : Google Scholar : PubMed/NCBI

28 

Morgan DO: Cyclin-dependent kinases: Engines, clocks, and microprocessors. Annu Rev Cell Dev Biol. 13:261–291. 1997. View Article : Google Scholar : PubMed/NCBI

29 

Lew DJ, Dulić V and Reed SI: Isolation of three novel human cyclins by rescue of G1 cyclin (Cln) function in yeast. Cell. 66:1197–1206. 1991. View Article : Google Scholar : PubMed/NCBI

30 

Bates S, Bonetta L, MacAllan D, Parry D, Holder A, Dickson C and Peters G: CDK6 (PLSTIRE) and CDK4 (PSK-J3) are a distinct subset of the cyclin-dependent kinases that associate with cyclin D1. Oncogene. 9:71–79. 1994.PubMed/NCBI

31 

Koff A, Giordano A, Desai D, Yamashita K, Harper JW, Elledge S, Nishimoto T, Morgan DO, Franza BR and Roberts JM: Formation and activation of a cyclin E-cdk2 complex during the G1 phase of the human cell cycle. Science. 257:1689–1694. 1992. View Article : Google Scholar : PubMed/NCBI

32 

Weinberg RA: The retinoblastoma protein and cell cycle control. Cell. 81:323–330. 1995. View Article : Google Scholar : PubMed/NCBI

33 

Ben-Neriah Y and Karin M: Inflammation meets cancer, with NF-κB as the matchmaker. Nat Immunol. 12:715–723. 2011. View Article : Google Scholar : PubMed/NCBI

34 

Jost PJ and Ruland J: Aberrant NF-kappaB signaling in lymphoma: Mechanisms, consequences, and therapeutic implications. Blood. 109:2700–2707. 2007.

35 

Viatour P, Merville MP, Bours V and Chariot A: Phosphorylation of NF-kappaB and IkappaB proteins: Implications in cancer and inflammation. Trends Biochem Sci. 30:43–52. 2005. View Article : Google Scholar : PubMed/NCBI

36 

Hess J, Angel P and Schorpp-Kistner M: AP-1 subunits: Quarrel and harmony among siblings. J Cell Sci. 117:5965–5973. 2004. View Article : Google Scholar : PubMed/NCBI

37 

Pahl HL: Activators and target genes of Rel/NF-kappaB transcription factors. Oncogene. 18:6853–6866. 1999. View Article : Google Scholar : PubMed/NCBI

38 

Drysdale MJ, Brough PA, Massey A, Jensen MR and Schoepfer J: Targeting Hsp90 for the treatment of cancer. Curr Opin Drug Discov Devel. 9:483–495. 2006.PubMed/NCBI

Related Articles

Journal Cover

August-2017
Volume 51 Issue 2

Print ISSN: 1019-6439
Online ISSN:1791-2423

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Ishikawa C, Senba M and Mori N: Butein inhibits NF-κB, AP-1 and Akt activation in adult T-cell leukemia/lymphoma. Int J Oncol 51: 633-643, 2017
APA
Ishikawa, C., Senba, M., & Mori, N. (2017). Butein inhibits NF-κB, AP-1 and Akt activation in adult T-cell leukemia/lymphoma. International Journal of Oncology, 51, 633-643. https://doi.org/10.3892/ijo.2017.4026
MLA
Ishikawa, C., Senba, M., Mori, N."Butein inhibits NF-κB, AP-1 and Akt activation in adult T-cell leukemia/lymphoma". International Journal of Oncology 51.2 (2017): 633-643.
Chicago
Ishikawa, C., Senba, M., Mori, N."Butein inhibits NF-κB, AP-1 and Akt activation in adult T-cell leukemia/lymphoma". International Journal of Oncology 51, no. 2 (2017): 633-643. https://doi.org/10.3892/ijo.2017.4026