Harmol induces autophagy and subsequent apoptosis in U251MG human glioma cells through the downregulation of survivin

  • Authors:
    • Akihisa Abe
    • Hiroko Kokuba
  • View Affiliations

  • Published online on: January 18, 2013     https://doi.org/10.3892/or.2013.2242
  • Pages: 1333-1342
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

The β-carboline alkaloids are plant substances that exhibit a wide spectrum of neuropharmacological, psychopharma­cological and antitumor effects. In the present study, we found that harmol, a β-carboline alkaloid, induced autophagy and suppression of survivin expression, and subsequently induced apoptotic cell death in U251MG human glioma cells. Autophagy was induced within 12 h by treatment with harmol. When treated for over 36 h, however, apoptotic cell death was induced. Harmol treatment also reduced survivin protein expression. Small interfering RNA (siRNA)-mediated knockdown of survivin enhanced the harmol-induced apoptosis. Knockdown of survivin by siRNA also induced autophagy. Therefore, harmol-induced apoptosis is a result of the reduction in survivin protein expression. Treatment with 3-methyladenine (3-MA) in the presence of harmol did not affect the expression of survivin and diminished harmol-induced cell death. Treatment with chloroquine in the presence of harmol did not suppress the reduction of survivin expression and increased harmol-induced cell death. From these results, harmol-induced reduction of survivin expression was closely related to autophagy. It is assumed that when isolation membrane formation is inhibited by treatment with 3-MA, reduction of survivin protein expression and apoptotic cell death were not induced. However, when isolation membrane formation is started and an autophagosome is formed, survivin expression is suppressed and apoptosis is executed. Harmol treatment reduced phosphorylation of Akt, mammalian target of rapamycin (mTOR) and its downstream targets p70-ribosomal protein S6 kinase and 4E-binding protein 1, resulting in induction of autophagy. Conversely, activation of the Akt/mTOR pathway inhibited harmol-induced autophagy and cell death. These findings indicate that harmol-induced autophagy involves the Akt/mTOR pathway. Taken together, autophagy induced by harmol represented a pro-apoptotic mechanism, and harmol suppressed the expression of survivin and subsequently induced apoptosis.

Introduction

Gliomas are the most common type of malignant brain tumors and are resistant to many types of treatments, including chemotherapy, radiation and other adjuvant therapies. Patients with the most malignant histopathologic subtype, glioblastoma, present with the worst prognosis, with a median survival length of less than one year, despite aggressive surgery with adjuvant radiotherapy and chemotherapy (1,2). In addition, glioma cells are prone to acquire drug resistance (3). Currently, there is still a need to identify chemotherapeutic agents with cytotoxic effects exclusively targeted against malignant glioma cells. Improved chemotherapeutic regimens and other strategies are urgently needed.

‘Apoptosis’, signifying genetically controlled programmed cell death (PCD), not only plays a crucial role during tissue development and homeostasis, but is also involved in a wide range of pathologies (4). Since the 1960s, various morphological forms of PCD have been recognized. Clarke classified cell death into 4 types, including type I PCD (apoptosis) and type II PCD (autophagy) (5). Evasion of apoptosis is a hallmark of most malignant cells and contributes to the insensitivity to various current cancer therapies (6). Numerous studies have demonstrated that most chemotherapeutic agents and certain naturally occurring compounds induce cell death by activating the apoptotic pathways. It is thought that apoptosis induction in tumor cells, with either drugs or natural products, is an effective therapy for cancer and immune system diseases.

Autophagy is one of the major regulatory mechanisms in the degradation of intracellular proteins and organelles (7,8). It is a highly conserved evolutionary process which occurs in the cells of all eukaryotic organisms, from yeasts to humans. During autophagy, the cytosol and entire organelles become encased in double-membrane-bound vacuoles (autophagosomes), and subsequently fuse with lysosomes to form autolysosomes and are eventually degraded by lysosomal hydrolases (7). Although originally characterized as a survival response to nutrient deficiency, autophagy is now recognized as being frequently induced in response to a variety of stressors to maintain cellular homeostasis (911). The importance of autophagy has been emphasized in various biological fields, including cancer (10,1214). Furthermore, cancer cells undergo less autophagy than normal cells (15,16). These findings indicate that autophagy induction is an attractive modality of anticancer therapy.

β-carboline alkaloids are present in several medicinal plants, including Peganum harmala, Passiflora incarnate, and Bansteriopsis caapi(17,18). These plants have been used in traditional medicine to treat asthma, jaundice, lumbago, and other human ailments (1820). The β-carboline alkaloids are also found in common plants (e.g., wheat, rice, soybeans, grapes) and plant-derived drinks (e.g., wine, beer, whisky, brandy) (21). It is known that they also exist in mammalian tissues (22,23). It has been reported that certain β-carboline alkaloids and their related compounds exhibit cytotoxic effects on cancer cells (2426). We also previously reported that harmol, a β-carboline alkaloid, induced apoptosis (27) and autophagic cell death in human non-small cell lung cancer cells (28).

In the present study, we investigated the anticancer effect of harmol on the U251MG human glioma cell line. Furthermore, we examined the types of cell death which were induced by harmol and the possible mechanisms.

Materials and methods

Chemicals

Harmol (purity, minimum 98%) (Fig. 1), dimethyl sulfoxide (DMSO), 3-methyladenine (3-MA) and chloroquine (CQ) were purchased from Wako Pure Chemical Industries, Ltd. (Osaka, Japan). Harmol was dissolved in DMSO at a concentration of 200 mM, and stored at −20°C until use. Monodansylcadaverine (MDC) was obtained from Sigma-Aldrich (St. Louis, MO, USA). Caspase substrates Ac-DEVD-7-amino-4-trifluoromethyl coumarin (AFC) (caspase-3), Ac-IETD-AFC (caspase-8), and Ac-LEHD-AFC (caspase-9) were obtained from MBL (Nagoya, Japan). A recombinant full-length human active Akt1 protein (rAkt1) was purchased from R&D Systems, Inc. (Minneapolis, MN, USA).

Cell lines and culture conditions

Human glioma cell line U251MG was obtained from the Human Science Research Resources Bank (Osaka, Japan). This cell line was cultured in Dulbecco’s modified Eagle’s medium (Wako) with 80 mg/l kanamycin sulfate (Wako) and heat-inactivated fetal bovine serum (Biowest, Miami, FL, USA) and maintained at 37°C in an incubator containing 95% air and 5% CO2.

Cell viability

Cell viability was assessed with the CellTiter-Blue cell viability assay kit (Promega Corp., Madison, WI, USA). In viable cells, resazurin, which is contained in the CellTiter-Blue cell viability assay reagent, is metabolized to the fluorescent product resorufin. The tumor cells (4.5×104 cells/well) were precultured in a 48-well flat-bottom microtiter plate overnight at 37°C in a 5% CO2 humidified chamber. Then, various concentrations of harmol were added and the cells were incubated for 24–48 h. After incubation, 100 μl of CellTiter-Blue assay reagent (Promega Corp) was added to each well and the cells were further incubated for 1.5 h. After incubation, the medium of each well was analyzed using a Powerscan HT microplate reader (DS Pharma Biomedical Co., Osaka, Japan) at an excitation wavelength of 560 nm and an emission wavelength of 590 nm. Cell viability was determined based on the fluorescence intensity of non-treated cells.

Assay for caspase activity

Caspase activity was measured using fluorogenic peptide substrates. Both untreated control cells and cells which had been treated with 75 μM harmol were washed with ice-cold phosphate-buffered saline (PBS) and suspended in lysis buffer [100 mM HEPES (pH 7.5), 150 mM NaCl, 1% Nonidet P-40, 1 mM EDTA, 1 mM dithiothreitol (DTT), 1 mM phenylmethylsulfonyl fluoride, 10 μM leupeptin and 1 μM pepstatin] for 20 min on ice, followed by centrifugation at 12000 × g for 10 min at 4°C. In total, 50–70 μg of protein in 40 μl of buffer solution was mixed with 10 μl of 5 mM fluorogenic report substrate, for each individual caspase: Ac-DEVD-AFC for caspase-3, Ac-IETD-AFC for caspase-8 and Ac-LEHD-AFC for caspase-9. The reaction mixture was added to 50 μl of assay buffer [20 mM HEPES (pH 7.4), 0.1 M NaCl, 5 mM DTT, 0.1% Nonidet P-40], and incubated at 37°C for 1 h. The enzymatic product, AFC, which was released from the substrate, was excited at 400 nm to measure its emission at 505 nm. Untreated cells were used as a control.

Visualization of MDC-labeled vacuoles in harmol-treated A549 cells

U251MG cells were treated with 100 μM harmol for 0–12 h at 37°C, and then further treated with 30 μM MDC for 10 min. After treatment, cells were washed 3 times with PBS and immediately observed by fluorescence microscopy (BZ-8100, Keyence Co., Osaka, Japan).

Electron microscopy

Cultured U251MG cells with and without 100 μM harmol treatment were rinsed with phosphate buffer (PB) (0.1 M, pH 7.4). Cells were then fixed with 2.5% glutaraldehyde in PB for 3 h at 4°C, rinsed with PBS for 5 min at least 3 times, and post-fixed with 1% osmium tetroxide for 1 h at 4°C, followed by dehydration in a graded ethanol series for 5 min at room temperature (RT). After dehydration with absolute ethanol, cells were treated with a mixture of absolute ethanol and Spurr resin (1:1, v/v) for 1 h at RT. Then, the mixture was replaced with pure Spurr resin and cells were treated for 1 h at RT, twice. The pellets were transferred to a BEEM capsule, overlaid with new resin and subjected to resin polymerization at 70°C for 16 h. Ultrathin sections were cut with an Ultracut E microtome (Reichert-Jung Co., Vienna, Austria), stained with uranyl acetate and lead citrate, and examined using a JEM-1230 transmission electron microscope (JEOL, Tokyo, Japan).

Western blot analysis

Whole cell proteins were isolated from untreated and harmol-treated U251MG cells. After treatment, the cells were centrifuged at 300 × g for 5 min, and the pellet was lysed in a buffer containing 25 mM HEPES (pH 7.4), 150 mM NaCl, 0.5% Triton X-100, 10% glycerol, 1 mM DTT, 1 mM sodium orthovanadate, 25 mM β-glycerophosphate, 1 mM NaF and 5 μl/ml protease inhibitor cocktail (Wako). The protein content of each lysate was determined using a BCA protein assay kit (Pierce Biotechnology Inc., Rockford, IL, USA). Protein lysates were then mixed with an equal volume of gel loading buffer [20% glycerol, 4% sodium dodecyl sulfate (SDS), 100 mM Tris, 5% β-mercaptoethanol and 0.01% bromophenol blue] before being boiled for 5 min. After boiling, 40 μg of protein was subjected to SDS-polyacrylamide gel electrophoresis. Proteins were then transferred onto a polyvinylidene fluoride membrane (PVDF; GE Healthcare UK Ltd., Buckinghamshire, UK). Blots were then blocked for 1 h at RT in 5% non-fat dry milk diluted in Tris-buffered saline supplemented with 0.1% Tween-20 (TBS-T) (ICN Biomedicals Inc., Aurora, OH, USA). The following primary antibodies were incubated overnight at RT in TBS-T as follows: rabbit anti-poly-(ADP-ribose)-polymerase (PARP) (Cell Signaling Technology Inc., Danvers, MA, USA; 1:1000), rabbit anti-Akt (Cell Signaling Technology; 1:1000), rabbit anti-phospho-Akt (p-Akt) (Cell Signaling Technology; 1:1000), rabbit anti-mammalian target of rapamycin (mTOR) (Cell Signaling Technology; 1:1000), rabbit anti-phospho-mTOR (p-mTOR) (Cell Signaling Technology; 1:1000), rabbit anti-70 kDa ribosomal protein S6 kinase (P70S6K) (Cell Signaling Technology; 1:1000), rabbit anti-phospho-P70S6K (p-P70S6K) (Cell Signaling Technology; 1:1000), rabbit anti-eukaryotic translation initiation factor 4E-binding protein 1 (4E-BP1) (Cell Signaling Technology; 1:1000), rabbit anti-phospho-4E-BP1 (p-4E-BP1) (Cell Signaling Technology; 1:1000), mouse anti-microtubule-associated protein 1 light chain 3 (LC3) (MBL; 1:1000), and mouse anti-β-actin (Sigma-Aldrich; 1:3000). Peroxidase-conjugated secondary antibodies were incubated for 1 h at RT as follows: goat anti-mouse IgG (Jackson Immunoresearch Laboratories Inc., West Grove, PA, USA; 1:5000) or goat anti-rabbit IgG (Cell Signaling Technology; 1:2000). Blots were then washed with TBS-T and developed using Immobilon chemiluminescent substrate (Millipore Corp., Billerica, MA, USA).

Survivin siRNA transfection

Knockdown of survivin expression in U251MG cells was achieved by transfection of siRNA. The control and survivin siRNAs (Cell Signaling Technology) were diluted to a final concentration of 20 nM in Opti-Mem I (Invitrogen Corp., Carlsbad, CA, USA), and transfection was performed with cells at 40–50% confluency using Lipofectamine 2000 transfection reagent (Invitrogen) according to the manufacturer’s protocol.

Statistical analysis

Results are presented as means ± SD. Statistical comparisons of the results were carried out using analysis of variance. A P-value of <0.05 was considered to indicate a statistically significant difference. Differences between the means of control and treated cells were analyzed by Dunnett’s test.

Results

Cytotoxic effects of harmol on U251MG cells

Continuous treatment of U251MG human glioma cells with harmol for 24 and 48 h induced growth inhibition and cell death in a time- and dose-dependent manner (Fig. 2). The cell viabilities of U251MG cells exposed to harmol at various concentrations of 12.5, 25, 50, 75, and 100 μM in triplicate for 24 h were 102.3, 99.8, 93.7, 82.3, and 66.7%, respectively. After treatment with harmol for 48 h, the cell viabilities decreased to 98.0, 93.6, 72.5, 38.2, and 17.3%, respectively (Fig. 2).

Changes in caspase activities and poly-(ADP-ribose)-polymerase (PARP) cleavage

To confirm whether the cell death induced by harmol treatment was due to apoptosis, the activities of 3 types of caspases were examined (Fig. 3A). Harmol showed no noticeable effect on the activities of these caspases within the 24-h treatment. However, in the 36-h treatment, activities of caspase-3 and -9 were elevated to 486 and 346%, respectively, while caspase-8 activity showed no significant change. We also examined the effects of harmol on PARP cleavage. PARP is a key participant in DNA base excision repair and in maintaining genome integrity. PARP is well known as an endogenous substrate for caspase-3 and an early marker of apoptosis (29). After treatment with harmol for 6–48 h, the cleaved form of PARP (89 kDa) was detected at 36 and 48 h following treatment of the cell lysate (Fig. 3B). These data indicated that treatment with harmol for over 36 h induced apoptosis in U251MG cells.

Harmol treatment suppresses survivin protein expression

While harmol was shown to induce apoptosis in U251MG cells, the mechanisms of inducing this phenomenon were unclear. Afterwards, we investigated the effect of harmol treatment on the expression of survivin protein, which is a known anti-apoptotic protein. From the results, harmol treatment was shown to suppress the expression of survivin protein in a dose- and time-dependent manner (Fig. 4).

Formation of autophagic vacuoles following treatment with harmol

Phase contrast microscopy of harmol-treated U251MG cells showed numerous, high-density vacuoles (Fig. 5, left lower panel). Based on the above findings, we would have expected these vacuoles to be formed upon harmol treatment; they were not detected in untreated cell autophagosomes. The fluorescent compound MDC is a specific marker for autophagosomes (30), and is commonly used to stain autophagic vesicles. We studied the incorporation of MDC in cells where autophagy was stimulated by harmol treatment. U251MG cells were treated with 100 μM harmol for 12 h and then analyzed by fluorescence microscopy. As shown in Fig. 5 (right upper panel), in control cells (0 h treatment), MDC-labeled vacuoles were scarcely detected. On the other hand, in cells which were treated with harmol for 12 h, numerous MDC-labeled fluorescent dots were clearly detected (Fig. 5, right lower panel). Although MDC-labeled fluorescent dots were observed in harmol-treated U251MG cells, in order to identify them more precisely, we performed electron microscopy to obtain ultrastructural information regarding the morphology of harmol-induced autophagy in U251MG cells. Fig. 6 shows an electron micrograph of a non-treated U251MG cell (left panel) and a U251MG cell treated with harmol (right panel). Although the electron micrograph of the harmol-treated U251MG cell showed numerous cytoplasmic phagolysosomes that contained highly electron-dense materials (Fig. 6, right panel), few phagolysosomes were observed in the control cell (Fig. 6, left panel).

Quantitative detection of microtubule-associated protein 1 light chain 3 (LC3)

To quantify the incidence of harmol-induced autophagy, we examined the expression of LC3-I and LC3-II proteins using western blot analysis, as these proteins are essential to the formation of autophagosomes (especially LC3-II) and have been widely used for estimating the number of autophagosomes or the incidence of autophagy (31). The expression of LC3-II protein in harmol-treated cells was detected within 12 h, and increased during treatment in a time-dependent manner (Fig. 7). Taken together, these results also indicated that harmol induced autophagy, and that harmol-induced autophagy preceded apoptosis in U251MG cells (Fig. 3 vs. Figs. 57).

Role of autophagy in harmol-mediated cell death

Autophagy plays an important role in the regulation of cell survival or death (7,8). To determine whether the induction of autophagy by harmol is related to survival or death mechanisms, we investigated the effect of harmol on cell survival in the presence of autophagy inhibitors in U251MG cells. As shown in Fig. 8A, 3-MA, a specific inhibitor of the early stage autophagic process (32), reduced the harmol-induced cell death in U251MG cells. On the other hand, the late stage autophagy inhibitor CQ (33) increased the harmol-induced cell death. On western blot analysis, although treatment with harmol alone induced the expression of LC3-II protein (known as an autophagy marker), co-treatment with harmol and 3-MA notably suppressed LC3-II protein expression (Fig. 8B). In contrast, co-treatment with harmol and CQ induced progressive accumulation of LC3-II throughout the 24 h duration assay (Fig. 8B). Following treatment with harmol alone for 24 h, although PARP cleavage was not induced (Figs. 3 and 8B), CQ treatment in the presence of harmol induced PARP cleavage. Regarding the anti-apoptotic protein survivin, although >24 h of treatment with 100 μM harmol strongly suppressed the expression of survivin protein (Fig. 6), the combination treatment of harmol with 3-MA did not suppress the expression of survivin protein (Fig. 8B). However, co-treatment with harmol and CQ did not suppress the reduction of survivin protein expression in U251MG cells (Fig. 8B).

Genetic suppression of survivin protein expression increases harmol-induced apoptosis

To demonstrate whether harmol-induced apoptosis is related to the suppression of survivin protein expression, we performed siRNA-mediated knockdown of survivin in U251MG cells. As a result, although in control-siRNA transfected U251MG cells, PARP cleavage was not detected within the 24 h treatment of harmol, in cells where the expression of survivin protein was supressed by transfection with survivin siRNA, PARP cleavage was detected within 24 h following treatment with harmol (Fig. 9B). In addition, siRNA depletion of survivin protein induced autophagy, which was confirmed by LC3-II protein expression (Fig. 9B, the non-treated control of the survivin-siRNA transfected cell sample was compared with the non-treated control of the control-siRNA transfected cell sample). Furthermore, in the cytotoxicity experiment, the suppression of survivin expression by siRNA increased the harmol-induced cytotoxicity (Fig. 9A).

Akt/mTOR signaling pathway is involved in harmol-induced autophagy in U251MG cells

Our next objective was to identify and characterize the molecular pathways involved in harmol-induced autophagy. The classical pathway that regulates autophagy involves the serine/threonine kinase, mTOR (34). The Akt/mTOR/P70S6K, 4E-BP1 pathway is the main regulatory pathway that negatively regulates autophagy. However, several pathways appear to regulate autophagy in mammalian cells. Thus, we investigated whether the autophagy induced by harmol is mediated by an mTOR-dependent or -independent pathway using western blot analysis (Fig. 10). Harmol treatment caused a significant time-dependent decrease in the phosphorylation of Akt protein in U251MG cells. Equally, harmol treatment diminished the levels of the phosphorylated (activated) form of mTOR, a downstream target of Akt which may inhibit cell growth and induce autophagy. Furthermore, examination of the phosphorylation of 2 downstream effectors of mTOR signaling, P70S6K and 4E-BP1, showed that both p-P70S6K and p-4E-BP1 decreased in a time-dependent manner, revealing a potent inhibitory effect of harmol treatment on the Akt/mTOR signaling pathway.

Activation of the Akt pathway inhibits harmol-induced autophagy and cytotoxicity in U251MG cells

It has been reported that the Akt/mTOR pathway mediates autophagy induced by various anticancer therapies and natural products (3537). Thus, we used rAkt1 to activate the Akt pathway as described previously (38). In Akt pathway-activated U251MG cells treated with rAkt1, we examined the effect on harmol-induced autophagy and cell death. U251MG cells were treated with 100 μM harmol, 500 ng/ml rAkt1, or both, for 12 h for western blotting. The addition of rAkt1 inhibited the harmol-induced decrease in phosphorylated Akt, mTOR and P70S6K in U251MG cells (Fig. 11A). Furthermore, the addition of rAkt1 significantly decreased harmol-induced autophagy, confirmed by LC3-II protein expression (Fig. 11A). In the cytotoxicity experiment, the addition of rAkt1 significantly suppressed harmol-induced cytotoxicity in U251MG cells (Fig. 11B).

Discussion

Previous studies have demonstrated that harmol induced apoptosis in human lung H596 cells through the caspase-8-dependent pathway, but independently of Fas/Fas ligand interaction (27). Furthermore, harmol induced autophagic cell death in human lung A549 cells (28). In the present study, we found that harmol induced autophagy and apoptotic cell death in U251MG cells.

Defects in the regulation of apoptosis are common phenomena in many types of cancer and are also the critical steps in tumorigenesis and resistance to therapy (39). There are two key signaling mechanisms through which cells may be triggered to undergo apoptosis. In the intrinsic pathway, diverse stimuli, such as reactive oxygen species, DNA-damaging reagents, and Ca2+ mobilizing agents provoke cell stress or damage, typically by activating one or more members of the BH3-only protein family, leading to the release of cytochrome c. Cytochrome c then binds to apoptosis-activating factor 1 (Apaf-1) and procaspase-9, resulting in the activation of caspase-9 by proteolytic cleavage. The extrinsic pathway starts with death receptor ligation or Fas/Fas ligand interaction, followed by oligomerization of the receptor, use of Fas-associated death domain (FADD) protein, and activation of caspase-8 (40). Caspase-8 and -9 have been generally defined as initiator caspases, and can in turn activate caspase-3, the executor of apoptosis (41,42). As shown in Fig. 3, over 36 h of treatment with harmol increased the activity of both caspase-9 and -3, and induced PARP cleavage. However, no noticeable change in the activity of caspase-8 was observed. From these results, it is assumed that harmol-induced apoptosis was associated with an intrinsic pathway instead of an extrinsic pathway, such as Fas/Fas ligand interaction or the TNF receptor pathway. Although the reasons for the differences in the responses to harmol treatment in U251MG, H596 and A549 cells are unclear, these phenomena in each of these cells might be based on the characteristics of each cell line and the involved stimuli.

Survivin is the smallest member of the inhibitor of apoptosis protein (IAP) family that is selectively overexpressed in most common types of human cancers, but not in normal adult tissues, and has been implicated in the control of cell division, inhibition of apoptosis, and tumor cell resistance to certain anticancer agents and ionizing radiation (4348). Therefore, these findings make survivin an attractive anticancer drug target. In U251MG cells, we found that harmol suppressed the expression of survivin protein and induced apoptosis associated with activation of caspases and cleavage of PARP, consistent with the role of survivin in antagonizing caspases, thereby blocking apoptosis. Although treatment with harmol alone induced cell death following a 36-h treatment (Fig. 3), the knockdown of survivin expression by specific siRNA sensitized U251MG cells to harmol-induced cell death. Cleaved PARP, a hallmark of apoptosis, was detected within 24 h (Fig. 9B). From these results, the possibility remains that harmol-induced apoptosis is due to the downregulation of survivin protein expression.

Of note, although harmol treatment induced not only apoptosis, but also autophagy (Figs. 57), the suppression of survivin protein expression by siRNA also induced autophagy (Fig. 9B, comparing the LC3-II band of the non-treated control sample of control siRNA with the LC3-II band of the non-treated control sample of survivin siRNA). Furthermore, survivin siRNA treatment increased harmol-induced autophagy, as evidenced by conversion of the autophagosomal marker LC3 protein from the cytosolic form (LC3-I) to the membrane-bound, phosphatidylethanolamine-conjugated form (LC3-II), associated with a size shift detected by western blotting (Fig. 9B). In brief, it appears that at first, harmol induces autophagy (induced within 12 h by treatment with harmol) (Fig. 7) and subsequently reduces the expression of survivin protein. Next, the reduction of survivin protein expression increases autophagy, and finally, apoptosis is induced. Autophagy plays a critical role in the regulation of cell survival or death (12). Therefore, to investigate whether the induction of autophagy by harmol has a cytoprotective function or leads to cell death (i.e., autophagic cell death), we treated U251MG cells with harmol and measured the effects on cell survival or apoptosis in the presence or absence of pharmacological autophagy inhibitors. Although 3-MA treatment significantly reduced harmol-induced cell death, CQ treatment highly enhanced harmol-induced cell death (Fig. 8). In addition, although harmol treatment decreased the expression of survivin, 3-MA treatment in the presence of harmol suppressed the reduction of survivin expression (Fig. 8B). On the other hand, CQ treatment in the presence of harmol did not suppress the reduction of survivin expression (Fig. 8B). The 3-MA molecule, a specific inhibitor of the early-stage autophagic process, is known as a PI3K inhibitor. Class III phosphatidylinositol-3-kinase (class III PI3K) is closely associated with autophagosome membrane formation (10). Therefore, treatment with 3-MA inhibits the formation of autophagic vacuoles. On the other hand, the late-stage autophagy inhibitor CQ blocks the fusion of autophagosomes and lysosomes, thereby suppressing the degradation of the contents within the autophagosome. Considering this information, the possibility remains that when harmol-induced autophagosome formation is inhibited, suppression of survivin protein expression is not induced. In contrast, if an autophagosome is formed, suppression of survivin protein expression occurs, and subsequently apoptosis is induced. In brief, it is assumed that harmol-induced autophagy plays pro-apoptotic roles. However, the detailed mechanisms of these phenomena are unclear. Additional experiments are required to clarify these mechanisms.

Autophagy is regulated by multiple signaling pathways as diverse as class III PI3K (10), protein kinase mTOR (7,49,50), mitogen activated protein kinase (MAPK) (51,52), AMP-activated protein kinase (AMPK) (53,54), Beclin1 (55) and bcl-2 (56). In particular, Akt/mTOR signaling is the major pathway and plays a variety of physiologic roles, including cell growth, cell cycle regulation, migration and survival (5759). Furthermore, this pathway is also frequently associated with oncogenesis (60). It is known that the Akt/mTOR pathway negatively regulates autophagy (10). Several studies have indicated that inhibition of the Akt/mTOR pathway is associated with the triggering of autophagy in cancer cells (6063). In U251MG cells, harmol treatment inhibited the phosphorylation of Akt (both Ser473 and Thr308), and moreover, exposure of cells to harmol also inactivated mTOR and reduced phosphorylation of its downstream targets p70S6K and 4E-BP1 (Fig. 10). In contrast, the activation of the Akt pathway by rAkt1 treatment increased the phosphorylation of mTOR and its downstream targets p70S6K and 4E-BP1 (Fig. 11A). Treatment with rAkt1 also improved the viability of harmol-treated U251MG cells (Fig. 11B). These results are consistent with many studies indicating that inhibition of the Akt/mTOR pathway is associated with induction of autophagy in cancer cells (35,36,50,61).

β-carbolines which occur naturally as harmala alkaloids in Peganum harmala Linné, are also widely found in other medicinal plants (17,18), and are present endogenously in mammalian tissues (22,23). Harmala alkaloids have been used in certain hallucinogenic preparations by some South American and African tribes (62). Furthermore, plants containing harmala alkaloids have long been used in traditional medicine to treat asthma, jaundice, lumbago, and other ailments (1820). Moreover certain β-carboline alkaloids have a wide spectrum of neuropharmacological and psychopharmacological actions on the central nervous system, such as tremorogenesis (63,64), hypothermia (65), hallucinogenesis (66,67), monoamine oxidase inhibition (68,69), and convulsive or anticonvulsive actions (70). The β-carboline-induced central nervous effects mainly occur with β-carbolines which have a methoxyl group at the C-7 position in their structures (71). Therefore, β-carbolines which have a methoxyl group at the C-7 position are not suitable for chemotherapeutic agents. On the other hand, it is reported that harmol, which has a hydroxyl group at the C-7 position (Fig. 1), showed only slight central depression and convulsive effects (72). Furthermore, harmol has either slight or no tremor-inducing action (73,74).

With regard to its metabolic pathway, harmol is conjugated with sulfate and excreted in the urine and bile; it has a rapid rate of urinary excretion in humans after intravenous administration (75). Therefore, harmol is considered to have low toxicity in humans and animals.

In conclusion, we demonstrated that harmol induces autophagy and apoptosis in U251MG human glioma cells. As these effects of harmol were investigated only in glioma cells, harmol’s effects on other types of cancer should also be examined. Furthermore, although additional investigation is needed to precisely clarify the autophagy and apoptosis induction pathway, it is thought to be a candidate pathway that can be targeted by a chemotherapeutic agent in cancer treatment.

Acknowledgements

We are indebted to Dr Clifford A. Kolba and Associate Professor Edward F. Barroga of the Department of International Medical Communications of Tokyo Medical University for their editorial review of the English manuscript.

References

1 

DeAngelis LM: Brain tumors. N Engl J Med. 344:114–123. 2001. View Article : Google Scholar : PubMed/NCBI

2 

Phillips HS, Kharbanda S, Chen R, Forrest WF, Soriano RH, Wu TD, et al: Molecular subclasses of high-grade glioma predict prognosis, delineate a pattern of disease progression, and resemble stages in neurogenesis. Cancer Cell. 9:157–173. 2006. View Article : Google Scholar : PubMed/NCBI

3 

Kanzawa T, Kondo Y, Ito H, Kondo S and Germano I: Induction of autophagic cell death in malignant glioma cells by arsenic trioxide. Cancer Res. 63:2103–2108. 2003.PubMed/NCBI

4 

Fadeel B, Orrenius S and Zhivotovsky B: Apoptosis in human disease: a new skin for the old ceremony? Biochem Biophys Res Commun. 266:699–717. 1999. View Article : Google Scholar : PubMed/NCBI

5 

Clarke PG: Developmental cell death: morphological diversity and multiple mechanisms. Anat Embryol. 181:195–213. 1990. View Article : Google Scholar : PubMed/NCBI

6 

Brown JM and Attardi LD: The role of apoptosis in cancer development and treatment response. Nat Rev Cancer. 5:231–237. 2005.PubMed/NCBI

7 

Klionsky DJ and Emr SD: Autophagy as a regulated pathway of cellular degradation. Science. 290:1717–1721. 2000. View Article : Google Scholar : PubMed/NCBI

8 

Levine B and Klionsky DJ: Development by self-digestion: molecular mechanisms and biological functions of autophagy. Dev Cell. 6:463–477. 2004. View Article : Google Scholar : PubMed/NCBI

9 

Shintani T and Klionsky DJ: Autophagy in health and disease: a double-edged sword. Science. 306:990–995. 2004. View Article : Google Scholar : PubMed/NCBI

10 

Gozuacik D and Kimchi A: Autophagy as a cell death and tumor suppressor mechanism. Oncogene. 23:2891–2906. 2004. View Article : Google Scholar : PubMed/NCBI

11 

Debnath J, Baehrecke EH and Kroemer G: Does autophagy contribute to cell death? Autophagy. 1:66–74. 2005. View Article : Google Scholar : PubMed/NCBI

12 

Mizushima N, Levine B, Cuervo AM and Klionsky DJ: Autophagy fights disease through cellular self-digestion. Nature. 451:1069–1075. 2008. View Article : Google Scholar : PubMed/NCBI

13 

Ogier-Denis E and Codogno P: Autophagy: a barrier or an adaptive response to cancer. Biochim Biophys Acta. 1603:113–128. 2003.PubMed/NCBI

14 

Corcelle E, Nebout M, Bekri S, et al: Disruption of autophagy at the maturation step by the carcinogen lindane is associated with the sustained mitogen-activated protein kinase/extracellular signal-regulated kinase activity. Cancer Res. 66:6861–6870. 2006. View Article : Google Scholar

15 

Otsuka H and Moscowitz M: Differences in the rates of protein degradation in untransformed and transformed cell lines. Exp Cell Res. 112:127–135. 1978. View Article : Google Scholar : PubMed/NCBI

16 

Kirkegaad K, Taylor MP and Jacson WT: Cellular autophagy: surrender, avoidance and subversion by microorganisms. Nat Rev Microbiol. 2:301–314. 2004. View Article : Google Scholar : PubMed/NCBI

17 

List PH and Hörhammer L: Hager’s Hand book for Pharmaceutical Practice. Springer-Verlag; Berlin: 1970

18 

Naranjo C: Psychotropic properties of the harmala alkaloids. Ethnopharmacological Search for Psychoactive Drugs. Efron DH, Holmestedt B and Kline NS: U.S. Public Health Service; New York: pp. 385–392. 1967

19 

Nadikarni KM: Indian Material Medica. 1. Popular Pakistan Limited; Bombay: pp. 927–929. 1976

20 

Dymock W, Warden CJ and Hooper D: Pharmacopia Indica. 1. Harmad National Foundation of Pakistan; pp. 252–253. 1976

21 

Khan SI, Abourashed EA, Khan IA and Walker LA: Transport of human alkaloids across Caco-2 cell monolayers. Chem Pharm Bull. 52:394–397. 2004. View Article : Google Scholar : PubMed/NCBI

22 

Airaksinen MM and Kari I: β-carbolines, psychoactive compounds in the mammalian body. Part I: occurrence, origin and metabolism. Med Biol. 59:21–34. 1981.

23 

Beck O and Faull KF: Concentrations of the enantiomers of 5-hydroxymethylptoline in mammalian urine: Implications for in vivo biosynthesis. Biochem Pharmacol. 65:97–106. 1986.PubMed/NCBI

24 

Sobhani AM, Ebrahimi SA and Mahmoudian MJ: An in vitro evaluation of human DNA topoisomerase I inhibition by Peganum harmala L. seed extract and its beta-carboline alkaloids. Pharm Pharm Sci. 5:19–23. 2002.PubMed/NCBI

25 

Lamchouri F, Setaff A, Cherrah Y, et al: Antitumor principles from Peganum harmala seeds. Therapie. 54:753–758. 1999.PubMed/NCBI

26 

Kuo PC, Shi LS, Damu AG, et al: Cytotoxic and antimalarial beta-carboline alkaloids from the roots of Eurycoma longifolia. J Nat Prod. 66:1324–1327. 2003. View Article : Google Scholar : PubMed/NCBI

27 

Abe A and Yamada H: Harmol induces apoptosis by caspase-8 activation independently of Fas/Fas ligand interaction in human lung carcinoma H596 cells. Anticancer Drugs. 20:373–381. 2009. View Article : Google Scholar : PubMed/NCBI

28 

Abe A, Yamada H, Moriya S and Miyazawa K: The β-carboline alkaloid harmol induces cell death via autophagy but not apoptosis in human non-small cell lung cancer A549 cells. Biol Pharm Bull. 34:1264–1272. 2011.

29 

Oliver FJ, de la Rubia G, Rolli V, Ruiz-Ruiz MC, de Murcia G and Murcia JM: Importance of poly(ADP-ribose) polymerase and its cleavage in apoptosis. Lesson from an uncleavable mutant. J Biol Chem. 273:33533–33539. 1998. View Article : Google Scholar : PubMed/NCBI

30 

Biederbick A, Kern HF and Elsässer HP: Monodansylcadaverine (MDC) is a specific in vivo marker for autophagic vacuoles. Eur J Cell Biol. 66:3–14. 1995.PubMed/NCBI

31 

Kabeya Y, Mizusima N, Ueno T, et al: LC3, a mammalian homologue of yeast Apg8p, is localized in autophagosome membranes after processing. EMBO J. 19:5720–5728. 2000. View Article : Google Scholar : PubMed/NCBI

32 

Seglen PO and Gordon PB: 3-Methyladenine: specific inhibitor of autophagic/lysosomal protein degradation in isolated rat hepatocytes. Proc Natl Acad Sci USA. 79:1889–1892. 1982. View Article : Google Scholar : PubMed/NCBI

33 

Kawai A, Uchiyama H, Takano S, Nakamura S and Ohkuma S: Autophagosome-lysosome fusion depends on the pH in acidic compartments in CHO cells. Autophagy. 3:154–157. 2007. View Article : Google Scholar : PubMed/NCBI

34 

Rubinsztein DC, Gestwicki JE, Murphy LO and Klionsky DJ: Potential therapeutic applications of autophagy. Nat Rev Drug Discov. 6:304–312. 2007. View Article : Google Scholar : PubMed/NCBI

35 

Fu L, Kim AY, Wang X, et al: Perifosine inhibits mammalian target of rapamycin signaling through facilitating degradation of major components in the mTOR axis and induces autophagy. Cancer Res. 69:8967–8976. 2009. View Article : Google Scholar : PubMed/NCBI

36 

Viola G, Bortolozzi R, Hamel E, et al: MG-2477, a new tubulin inhibitor, induces autophagy through inhibition of the Akt/mTOR pathway and delayed apoptosis in A549 cells. Biochem Pharmacol. 8:16–26. 2012. View Article : Google Scholar : PubMed/NCBI

37 

Hung JY, Hsu YL, Li CT, et al: 6-Shogaol, an active constituent of dietary ginger, induces autophagy by inhibiting the AKT/mTOR pathway in human non-small cell lung cancer A549 cells. J Agric Food Chem. 57:9809–9816. 2009. View Article : Google Scholar : PubMed/NCBI

38 

Takeuchi H, Kanzawa T, Kondo Y and Kondo S: Inhibition of platelet-derived growth factor signaling induces autophagy in malignant glioma cells. Br J Cancer. 90:1069–1075. 2004. View Article : Google Scholar : PubMed/NCBI

39 

Hanahan D and Weinberg RA: The hallmarks of cancer. Cell. 100:57–70. 2000. View Article : Google Scholar

40 

Gross A, Yin XM, Wang K, et al: Caspase cleaved BID targets mitochondria and is required for cytochrome c release, while BCL-XL prevents this release but not tumor necrosis factor-R1/Fas death. J Biol Chem. 274:1156–1163. 1999. View Article : Google Scholar : PubMed/NCBI

41 

Zha J, Weiler S, Oh KJ, Wei MC and Korsmeyer SJ: Posttranslational N-myristoylation of BID as a molecular switch for targeting mitochondria and apoptosis. Science. 290:1761–1765. 2000. View Article : Google Scholar : PubMed/NCBI

42 

Shao RG, Cao CX, Nieves-Neira W, Dimanche-Boitrel MT, Solary E and Pommier Y: Activation of the Fas pathway independently of Fas ligand during apoptosis induced by camptothecin in p53 mutant human colon carcinoma cells. Oncogene. 20:1852–1859. 2001. View Article : Google Scholar : PubMed/NCBI

43 

Deveraux Q and Reed JC: IAP family proteins: suppressors of apoptosis. Genes Dev. 13:239–252. 1999. View Article : Google Scholar : PubMed/NCBI

44 

Li F, Ambrosini G, Chu EY, et al: Control of apoptosis and mitotic spindle checkpoint by survivin. Nature. 396:580–584. 1998. View Article : Google Scholar : PubMed/NCBI

45 

Monzo M, Rosell R, Felip E, et al: A novel anti-apoptosis gene: re-expression of surviving messenger RNA as a prognosis marker in non-small-cell lung cancers. J Clin Oncol. 17:2100–2104. 1999.PubMed/NCBI

46 

Adida C, Haioun C, Gaulard P, et al: Prognostic significance of surviving expression in diffuse large B-cell lymphomas. Blood. 96:1921–1925. 2000.PubMed/NCBI

47 

Kato J, Kuwabara Y, Mitani M, et al: Expression of surviving in esophageal cancer: correlation with the prognosis and response to chemotherapy. Int J Cancer. 95:92–95. 2001. View Article : Google Scholar : PubMed/NCBI

48 

Altieli DC: Survivin, cancer networks and pathway-directed drug discovery. Nat Rev Cancer. 8:61–70. 2008. View Article : Google Scholar : PubMed/NCBI

49 

Levine B and Yuan J: Autophagy in cell death: an innocent convict? J Clin Invest. 115:2679–2688. 2005. View Article : Google Scholar : PubMed/NCBI

50 

Guertin DA and Sabatini DM: Defining the role of mTOR in cancer. Cancer Cell. 12:9–22. 2007. View Article : Google Scholar

51 

Petiot A, Pattingre S, Arico S, Meley D and Codongo P: Diversity of signaling controls of macroautophagy in mammalian cells. Cell Struct Funct. 27:431–441. 2002. View Article : Google Scholar : PubMed/NCBI

52 

Pattingre S, Bauvy C and Codongo P: Amino acids interfere with the ERK1/2-dependent control of macroautophagy by controlling the activation of Raf-1 in human colon cancer HT-29 cells. J Biol Chem. 278:16667–16674. 2003. View Article : Google Scholar : PubMed/NCBI

53 

Meley D, Bauvy C, Houben-Weerts JH, et al: AMP-activated protein kinase and the regulation of autophagic proteolysis. J Biol Chem. 281:34870–34879. 2006. View Article : Google Scholar : PubMed/NCBI

54 

Herrero-Martín G, Høyer-Hansen M, García-García C, et al: TAK1 activates AMPK-dependent cytoprotective autophagy in TRAIL-treated epithelial cells. EMBO J. 28:677–685. 2009.PubMed/NCBI

55 

Djavaheri-Mergny M, Maiuri MC and Kroemer G: Cross talk between apoptosis and autophagy by caspase-mediated cleavage of Beclin 1. Oncogene. 29:1717–1719. 2010. View Article : Google Scholar : PubMed/NCBI

56 

Zhou F, Yang Y and Xing D: Bcl-2 and Bcl-xL play important roles in the crosstalk between autophagy and apoptosis. FEBS J. 278:403–413. 2011. View Article : Google Scholar : PubMed/NCBI

57 

Lin CC, Lee CW, Chu TH, et al: Transactivation of Src, PDGF receptor, and Akt is involved in IL-1beta-induced ICAM-1 expression in A549 cells. J Cell Physiol. 211:771–780. 2007. View Article : Google Scholar : PubMed/NCBI

58 

Russo P, Catassi A, Cesario A and Servent D: Development of novel therapeutic strategies for lung cancer: targeting the cholinergic system. Curr Med Chem. 13:3493–3512. 2006. View Article : Google Scholar : PubMed/NCBI

59 

Akca H, Tani M, Hishida T, Matsumoto S and Yokota J: Activation of the AKT and STAT3 pathways and prolonged survival by a mutant EGFR in human lung cancer cells. Lung Cancer. 54:25–33. 2006. View Article : Google Scholar : PubMed/NCBI

60 

Kishimoto H, Ohteki T, Yajima N, et al: The Pten/PI3K pathway governs the homeostasis of Valpha14iNKT cells. Blood. 109:3316–3324. 2007. View Article : Google Scholar : PubMed/NCBI

61 

Takeuchi H, Kondo Y, Fujiwara K, et al: Synergistic augmentation of rapamycin-induced autophagy in malignant glioma cells by phosphatidylinositol 3-kinase/protein kinase B inhibitors. Cancer Res. 65:3336–3346. 2005.PubMed/NCBI

62 

Rommelspacher H, Nanz C, Borbe HO, Fehske KJ, Müller WE and Wollert U: 1-Methyl-beta-carboline (harmane), a potent endogenous inhibitor of benzodiazepine receptor binding. Naunyn Schmiedebergs Arch Pharmacol. 314:97–100. 1980. View Article : Google Scholar : PubMed/NCBI

63 

Ho BT: Pharmacological and biochemical studies with beta-carboline analogs. Curr Dev Psychopharmacol. 4:151–177. 1977.PubMed/NCBI

64 

Poirier LJ, Sourkes TL, Bouvier G, Boucher R and Carabin S: Striatal amines, experimental tremor and the effect of harmaline in the monkey. Brain. 89:37–52. 1966. View Article : Google Scholar : PubMed/NCBI

65 

Bruinvels J and Sourkes TL: Influence of drugs on the temperature-lowering effect of harmaline. Eur J Pharmacol. 4:31–39. 1969. View Article : Google Scholar

66 

Zetler G, Back G and Iven H: Pharmacokinetics in the rat of the hallucinogenic alkaloids harmine and harmaline. Naunyn Shemiedebergs Arch Pharmacol. 285:273–292. 1974. View Article : Google Scholar : PubMed/NCBI

67 

Grella B, Dukat M, Young R, et al: Investigation of hallucinogenic and related beta-carbolines. Drug Alcohol Depend. 50:99–107. 1998. View Article : Google Scholar : PubMed/NCBI

68 

Udenfriend S, Wiycop B, Redfield BG and Weissbach H: Studies with reversible inhibitors of monoamine oxidase: harmaline and related compounds. Biochem Pharmacol. 1:160–165. 1958. View Article : Google Scholar

69 

McIsaac WM and Estevez V: Structure-action relationship of beta-carbolines as monoamine oxidase inhibitors. Biochem Pharmacol. 15:1625–1627. 1966. View Article : Google Scholar : PubMed/NCBI

70 

Loew GH, Nienow J, Lawson JA, Toll L and Uyeno ET: Theoretical structure-activity studies of beta-carboline analogs. Requirements for benzodiazepine receptor affinity and antagonist activity. Mol Pharmacol. 28:17–31. 1985.

71 

Kim H, Sablin SO and Ramsay RR: Inhibition of monoamine oxidase A by β-carboline derivatives. Arch Biochem Biophys. 337:137–142. 1997.

72 

Gunn JA: Relations between chemical constitution, pharmacological actions, and therapeutic uses in the harmine group of alkaloids. Arch Int Pharmacodyn. 50:379–396. 1935.

73 

Zetler G, Singbart G and Schlosser L: Cerebral pharmacokinetics of tremor-producing harmala and iboga alkaloids. Pharmacology. 7:237–248. 1972. View Article : Google Scholar : PubMed/NCBI

74 

Yu AM, Idle JR, Krausz KW, Küpfer A and Gonzalez FJ: Contribution of individual cytochrome P450 isozymes to the O-demethylation of the psychotropic beta-carboline alkaloids harmaline and harmine. J Pharmacol Exp Ther. 305:315–322. 2003. View Article : Google Scholar : PubMed/NCBI

75 

Slotkin TA, DiStefano V and Au WY: Blood levels and urinary excretion of harmine and its metabolites in man and rats. J Pharmacol Exp Ther. 173:26–30. 1970.PubMed/NCBI

Related Articles

Journal Cover

April 2013
Volume 29 Issue 4

Print ISSN: 1021-335X
Online ISSN:1791-2431

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Abe A and Abe A: Harmol induces autophagy and subsequent apoptosis in U251MG human glioma cells through the downregulation of survivin. Oncol Rep 29: 1333-1342, 2013
APA
Abe, A., & Abe, A. (2013). Harmol induces autophagy and subsequent apoptosis in U251MG human glioma cells through the downregulation of survivin. Oncology Reports, 29, 1333-1342. https://doi.org/10.3892/or.2013.2242
MLA
Abe, A., Kokuba, H."Harmol induces autophagy and subsequent apoptosis in U251MG human glioma cells through the downregulation of survivin". Oncology Reports 29.4 (2013): 1333-1342.
Chicago
Abe, A., Kokuba, H."Harmol induces autophagy and subsequent apoptosis in U251MG human glioma cells through the downregulation of survivin". Oncology Reports 29, no. 4 (2013): 1333-1342. https://doi.org/10.3892/or.2013.2242