Skip to main content
Erschienen in: Cancer Cell International 1/2016

Open Access 01.12.2016 | Primary research

Selected polyphenols potentiate the apoptotic efficacy of glycolytic inhibitors in human acute myeloid leukemia cell lines. Regulation by protein kinase activities

verfasst von: Elena de Blas, María Cristina Estañ, María del Carmen Gómez de Frutos, Javier Ramos, María del Carmen Boyano-Adánez, Patricio Aller

Erschienen in: Cancer Cell International | Ausgabe 1/2016

Abstract

Background

The glycolysis inhibitor 2-deoxy-d-glucose (2-DG) is a safe, potentially useful anti-tumour drug, but its efficacy is normally low when used alone. Recent studies indicated that 2-DG stimulates the PI3K/Akt and MEK/ERK defensive pathways, which limits the apoptotic efficacy in tumour cell lines. We hypothesized that co-treatment with selected polyphenols could improve 2-DG-provoked apoptosis by preventing defensive kinase activation.

Methods

Cell proliferation was measured by cell counting or the MTT assay. Cell cycle, apoptosis and necrosis were determined by propidium iodide staining and/or annexin V labeling followed by flow cytometry. Mitochondria pore transition and depolarization were determined by calcein-ATM or rhodamine 123 labeling followed flow cytometry. Intracellular reactive oxygen species and GSH were determined by dichlorodihydrofluorescein diacetate or monochlorobimane labeling followed by flow cytometry or fluorimetry. Expression and phosphorylation of protein kinases were analyzed by the Western blot.

Results

(i) 2-DG-provoked apoptosis was greatly potentiated by co-treatment with the sub-lethal concentrations of the flavonoid quercetin in human HL60 acute myeloblastic leukemia cells. Allowing for quantitative differences, apoptosis potentiation was also obtained using NB4 promyelocytic and THP-1 promonocytic cells, using curcumin or genistein instead of quercetin, and using lonidamine instead of 2-DG, but not when 2-DG was substituted by incubation in glucose-free medium. (ii) Quercetin and 2-DG rapidly elicited the opening of mitochondria pore transition, which preceded the trigger of apoptosis. (iii) Treatments did not affect GSH levels, and caused disparate effects on reactive oxygen species generation, which did not match the changes in lethality. (iv) 2-DG and lonidamine stimulated defensive Akt and ERK phosphorylation/activation, while glucose starvation was ineffective. Polyphenols prevented the stimulation of Akt phosphorylation, and in some cases also ERK phosphorylation. In addition, quercetin and 2-DG stimulated GSK-3α,β phosphorylation/inactivation, although with different isoform specificity. The use of pharmacologic inhibitors confirmed the importance of these kinase modifications for apoptosis.

Conclusions

The present in vitro observations suggest that co-treatment with low concentrations of selected polyphenols might represent a manner of improving the poor anti-tumour efficacy of some glycolytic inhibitors, and that apoptosis potentiation may be at least in part explained by the regulation of defensive protein kinase activities.
Hinweise

Electronic supplementary material

The online version of this article (doi:10.​1186/​s12935-016-0345-y) contains supplementary material, which is available to authorized users.
Abkürzungen
Akt
protein kinase B
AML
acute myeloid leukemia
BIRB 796
1-5-tert-butyl-2-p-tolyl-2H-pyrazol-3-yl)-3-[4-(2-morpholin-4-yl-ethoxy)naphthalen-1-yl] urea
BSO
d l-buthionine-S,R-sulfoximine
Calcein-AM
calcein O,O’-diacetate tetrakis(acetoxymethyl) ester
Cur
curcumin
2-DG
2-deoxy-d-glucose
ERK
extracellular signal-regulated kinase
FITC
fluorescein isothiocyanate
Gen
genistein
GSH
reduced glutathione
H2DCFDA
dichlorodihydrofluorescein diacetate
Lon
lonidamine
LY294002
2-(4-Morpholinyl)-8-phenyl-4H-1-benzopyran-4-one
MAPK
mitogen-activated protein kinase
MEK
mitogen-induced extracellular kinase/extracellular signal-regulated kinase
mIMP
mitochondrial inner membrane permeability
MTT
3(4,5-dimethyl-2-thiazolyl)-2,5-diphenyl-2H-tetrazolium bromide
PBS
phosphate buffered saline
PD98059
2′-amino-3′-methoxyflavone
PI3K
phosphatidylinositol 3-kinase
PI
propidium iodide
R123
rhodamine 123
ROS
reactive oxygen species
SB203580
4-(4-fluorophenyl)-2-(4-methylsulfinylphenyl)-5-(4-pyridyl)-1H-imidazole
SB216763
3(2,4-dochlorophenyl)-4-(1-methyl-1H-indol-3-yl)-1H-pyrrole-2,5-diione
Triciribine
5-dihydro-5-methyl-1-β-d-ribofuranosyl-1,4,5,6,8-pentaazaacenaphtylen-3-amine hydrate
Quer
quercetin
U0126
1,4-diamino-2,3-dicyano-1,4-bis(2-aminophenylthio)butadiene
z-VAD-fmk
Z-Val-Ala-Asp(OMe)-CH2F

Background

A characteristic of tumour cells is the profound re-organization of metabolic parameters in relation to their healthy counterparts, allowing them to obtain the macromolecular constituents required for their rapid de-regulated growth and also as alternative sources of energy [13]. One of the best known modifications is the increased dependence on glucose metabolism instead of oxidative phosphorylation, even under aerobic conditions (a property known as “aerobic glycolysis” or “Warburg’’ effect). This peculiarity made possible the development of glycolysis-targeting drugs as potential anti-cancer agents. This category includes, among others, the glucose inactive analog 2-deoxy-d-glucose (2-DG) [4], the indazole derivative lonidamine (Lon) [5], and the small alkylating drug 3-bromopyruvate (3-BrP) [6]. Allowing for the disparity in chemical structure and hence in biochemical and molecular effects, these drugs target critical enzymes in the glycolytic pathway, namely hexokinase II (HKII) in the case of 2-DG and Lon [5, 7], and glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and to a lower extent HKII in the case of 3-BrP [8]. While promissory results have been obtained in clinical assays [5, 6, 9, 10], the efficacy of these agents is different. Thus, 3-BrP is quite toxic per se, while 2-DG and Lon are well tolerated but poorly efficacious in monotherapy. Nonetheless, 2-DG and Lon may be useful as radio- and chemo-sensitizing agents, overcoming resistance and increasing cyto-reduction by conventional anti-tumour treatments [4, 5, 9]. Using combinatory assays with the anti-leukemic agent arsenic trioxide (Trisenox), we recently demonstrated that a common effect of the anti-glycolytic drugs is the stimulation (albeit with different kinetics and intensity) of Akt/mTOR and MEK/ERK defensive pathways in several human acute myeloid leukemia (AML) cell lines, and that this stimulation restrains the apoptotic efficacy of 2-DG and Lon when used as single agents [11, 12]. Akt and/or ERK activation by 2-DG was also observed in other tumour cell models [1315].
Polyphenols represent a large collection of molecules present in the plant kingdom. At the low doses attainable in the daily diet these compounds exert multiple protective functions (e.g., against cellular oxidation, inflammation, aging, tumour initiation…). On the other hand, at high albeit still pharmacologically attainable concentrations many polyphenols selectively induce apoptosis in tumour cells, and exhibit clinical efficacy either alone or in combination with conventional anti-cancer drugs [16]. While the multiplicity of biochemical actions makes impossible to unequivocally ascribe their anti-cancer action to a single mechanism, a frequent effect of polyphenols is the inhibition of the PI3K/Akt defensive pathway [17, 18]. For instance, the flavonoid quercetin (Quer) is the natural analog of 2-(4-Morpholinyl)-8-phenyl-4H-1-benzopyran-4-one (LY294002), a potent PI3K inhibitor commonly used in laboratory research [19]. We previously observed that prolonged treatment (24–48 h) with sub-lethal concentrations of Quer, curcumin (Cur) and genistein (Gen) reduced the constitutive Akt phosphorylation in U937 and HL60 AML cells [2022]. On this ground, we hypothesized the pre-treatment with polyphenols might prevent Akt activation and as a consequence improve the lethality of glycolytic inhibitors. With this hypothesis in mind, in the present work we analyze the capacity of Quer, Cur and Gen to cooperate with 2-DG and Lon to induce apoptosis in HL60 and other AML cell lines. The regulatory function of Akt and other protein kinases, as well as the potential importance of other factors such as mitochondrial dysfunction and oxidative stress, are examined.

Methods

Reagents and antibodies

All components for cell culture were obtained from Invitrogen, Inc. (Carlsbad, CA). Dichlorodihydrofluorescein diacetate (H2DCFDA) and monochlorobimane were obtained from Molecular Probes, Inc. (Eugene, OR). The kinase inhibitors 1,4-Diamino-2,3-dicyano-1,4-bis(2-aminophenylthio)butadiene (U0126), 2′-Amino-3′-methoxyflavone (PD98059), 2-(4-Morpholinyl)-8-phenyl-4H-1-benzopyran-4-one (LY294002), 5-Dihydro-5-methyl-1-β-d-ribofuranosyl-1,4,5,6,8-pentaazaacenaphtylen-3-amine hydrate (triciribine hydrate, Akt inhibitor V), 4-(4-Fluorophenyl)-2-(4-methylsulfinylphenyl)-5-(4-pyridyl)-1H-imidazole (SB203580), and the caspase inhibitor Z-Val-Ala-Asp(OMe)-CH2F (z-VAD-fmk), were obtained from Calbiochem (Darmstad, Germany), and 1-5-tert-Butyl-2-p-tolyl-2H-pyrazol-3-yl)-3-[4-(2-morpholin-4-yl-ethoxy)naphthalene-1-yl] urea (BIRB 796) from Shelleck (Houston, TX). Rabbit anti-human p44/42 MAP kinase, phospho-p44/p42 MAP kinase (Thr202/Tyr204), Akt, phospho-Akt (Ser473) (D9E) XP™, p38 MAP kinase, phospho-p38 MAP kinase (Thr180/Tyr182), phospho-GSK-3α/β (Ser21/9), and phospho-S6 ribosomal protein (Ser235/236) polyclonal antibodies, were obtained from Cell Signaling Technology Inc. (Danvers, MA). Mouse GSK-3α/β monoclonal antibody (0011-A) was obtained from Santa Cruz Biotechnology, Inc. (Santa Cruz, CA). Peroxidase-conjugated immunoglobulin G antibodies were from DAKO Diagnostics, S.A. (Barcelona, Spain). All other non-mentioned reagents and antibodies were from Sigma (Madrid, Spain).

Cells and treatments

HL60 myeloblastic cells [23] and THP-1 promononocytic cells [24] were obtained from our institutional repository (CIB), and NB4 promyelocytic cells [25] were kindly supplied by Profs. M.D. Delgado and J. León (Departamento de Biología Molecular, Facultad de Medicina, Universidad de Cantabria, Santander, Spain). These cell lines represent distinct subtypes of human AML cells (HL60, M2; NB4, M3; THP-1, M5, according to the classification of the French-American-British (FAB) cooperative group), with substantial differences in molecular and biochemical parameters, and hence in the capacity of response to anti-cancer agents. Absence of mycoplasma contamination, and authentication by STR analysis, specific antigen expression, and PML-RARα fusion protein expression (NB4 cells) were corroborated by us or our technical staff. Cell handling (and all experimental procedures in general) was carried out strictly following the regulations of the Bioethics and Biosafety Commission of our Institution (Centro de Investigaciones Biológicas, CSIC). Conditions of cell growth and treatment were described in detail in preceding publications [11, 12]. For glucose starvation (Glu−), cells were extensively washed with phosphate-buffered saline (PBS) and then seeded at the appropriate concentration in glucose-lacking RPMI medium supplemented with 10 % (v/v) serum. For good comparison, the corresponding controls were subject to the same manipulation, but finally seeded in complete medium (Glu+).
Calcein-AM was commercially obtained as a 4 mM solution in dimethyl sulfoxide. Rhodamine 123 (R123, 1 mg/ml) was prepared in ethanol. Stock solutions of Lon (100 mM), Quer (100 mM), Gen (50 mM), Cur (20 mM), H2DCFDA (5 mM), monochlorobimane (200 mM), U0126 (2.63 mM), PD98059, LY294002 and triciribine (20 mM each), SB203580 (13,2 mM), SB216763 (50 mM), BIRB 796 (0.1 mM) and z-VAD-fmk (25 mM) were prepared in dimethyl sulfoxide. 3(4,5-Dimethyl-2-thiazolyl)-2,5-diphenyl-2H-tetrazolium bromide (MTT) was dissolved at 5 mg/ml in PBS. All these solutions were stored at −20 °C. A stock solution of propidium iodide (PI, 1 mg/ml) was prepared in phosphate buffered saline (PBS), and stored at 4 °C. 2-DG and DL-buthionine-(S,R)-sulfoximine (BSO) were freshly prepared at 250 and 50 mM, respectively, in PBS. 3-Bromopyruvate was freshly prepared at 30 mM in PBS, and the pH of the solution was adjusted at 7.2 with NaOH.

Flow cytometry

The analysis of samples was carried out on an EPICS XL flow cytometer (Coulter, Hialeah, FL) equipped with an air-cooled argon laser tuned to 488 nm. The specific fluorescence signal corresponding to fluorescein isothiocyanate, H2DCFDA, calcein-AM and R123 was collected with a 525-nm band pass filter, and the signal corresponding to PI with a 620-nm band pass filter. A total of 104 cells were scored in each determination.

Cell proliferation, cell cycle, apoptosis and necrosis

Total cell proliferation was measured by cell counting of trypan-blue excluding cells, or by means of the MTT colorimetric assay. This later procedure gives an indirect estimation of the relative number of viable cells in the culture, based on changes in mitochondrial metabolic activity. Cell cycle phase distribution was routinely determined by cell permeabilization followed by PI staining and flow cytometry analysis. When convenient, the resulting histograms were analyzed with the FlowLogic program (Inivai, Victoria, Canada). This technique also provided an estimation of the frequency of apoptotic cells, characterized by low (sub-G1) DNA content. The criterion used for necrosis was the loss of plasma membrane integrity, as determined by free PI uptake into non-permeabilized cells and flow cytometry analysis. In addition, apoptosis and necrosis were determined simultaneously by double labeling with annexin V-FITC and PI followed by flow cytometry measurement using an Annexin V-FITC Apoptosis Detection Kit (Immunostep, Salamanca, Spain). This procedure allows the distinction between viable cells (annexin V-negative/PI-negative), early apoptotic cells (annexin V-positive/PI-negative), late apoptotic or necrotic cells (annexin V-positive/PI-positive), and genuine necrotic cells (annexin-V negative/PI-positive). Since the loss of plasma membrane integrity leading to free PI penetration is compatible with both genuine necrosis and late apoptosis (also termed “secondary” necrosis), the pan-caspase inhibitor z-VAD-fmk was occasionally used to discriminate between these two possibilities. A detailed description of all these techniques can be found in our preceding works [21, 26, and references therein].

Mitochondrial membrane permeabilization and membrane potential dissipation

Inner mitochondrial membrane permeabilization (mIMP) was determined using the calcein-AM/CoCl2 method, originally reported by Petronilli et al. [27]. Our adaptation for flow cytometry using HL60 cells was already described in a preceding article [11]. Mitochondrial membrane potential (ΔΨm) was determined using the cationic agent R123 and flow cytometry analysis, as previously described [26].

Reactive oxygen species and reduced glutathione levels

The intracellular accumulation of reactive oxygen species (ROS) was measured by flow cytometry using the ROS-sensitive probe H2DCFDA. The intracellular level of reduced glutathione (GSH) was measured in a Varioskan Flash microplate reader (Thermo Fisher Scientific Inc, Waltham, MA) at excitation wavelength of 390 nm and emission wavelength of 520 nm, using the fluorescent probe monochlorobimane. The detailed procedures were described in a previous publication [26].

Immunoblotting

Cells were collected by centrifugation, washed with PBS and total protein extracts were obtained by lysing them for 20 min at 4 °C in a buffer consisting of 20 mM Tris–HCl (pH 7.5) containing 137 mM NaCl, 2 mM EDTA, 10 % (v/v) glycerol, and 1 % Nonidet P-40, and supplemented with a protease inhibitor cocktail, 1 mM sodium orthovanadate, and 10 mM NaF. After brief sonication and centrifugation for 15 min at 10,000×g at 4 °C, the supernatants were collected, and samples containing equal amounts of proteins were resolved by SDS–polyacrylamide gel electrophoresis. The proteins were then transferred to polyvinylidene fluoride (PVDF) membranes and immunodetected, as previously described [28]. When convenient, the relative band intensities were quantified using the Quantity One 1-D Analysis Software, version 4.6 (Bio-Rad Laboratories, Inc., Hercules, CA).

Data analysis and presentation

Except when indicated, all experiments were repeated at least three times, and as a rule the results are expressed as the mean value ± SD. Statistical analyses were carried out using one way ANOVA with Dunnett or Bonferroni post-test, using SAS version 9.4 (SAS Institute, Cary NC). The Dunnett’s method was followed when comparing different treatments with controls, and Bonferroni’s when pairwise comparisons were performed. The symbols used were: &, to compare treatment vs. control; *, to compare pairs of single treatments; and #, to indicate that the value in a combined treatment is higher than the sum of values in the corresponding single treatments. Sum of values were obtained by considering single treatment as independent random variables. In all cases, single symbol means p < 0.05, double symbol p < 0.01, and triple symbol p < 0.001. n.s., non-significant.

Results

Cell proliferation and cell death

Firstly, we examined the capacity of Quer and 2-DG, alone and in combination, to affect proliferation activity and induce apoptosis at 24 h of treatment in HL60 cells. Because of the hypothesis advanced in the ‘‘Background’’ section, namely that polyphenols might prevent early regulatory gene responses elicited by metabolic inhibitors, in the combined treatments Quer was applied 2 h before 2-DG (and except when otherwise indicated, this procedure will be also followed in all experiments along the whole work). As shown in Fig. 1a, treatment with 5–20 μM Quer or 1–5 mM 2-DG separately caused a concentration-dependent decrease in viable cell number, as determined by the MTT assay, and the response was augmented when the drugs were used in combination. In spite of the evident effect on proliferation, treatment with either 5–20 µM Quer or 1-5 mM 2-DG separately caused very low (less than 10 %) apoptotic effect, as measured by the frequency of cells with sub-G1 DNA content in flow cytometry assays. Nonetheless, apoptosis was greatly potentiated in more than additive manner when the drugs were used in combination (Fig. 1b, c). On the ground of the obtained results, the concentrations of 20 µM Quer and 5 mM 2-DG were selected for the following experiments, except when otherwise indicated. The cooperative apoptotic action between Quer and 2-DG was confirmed using the annexin V/PI assay (Fig. 1d). Moreover, the pan-caspase inhibitor z-VAD-fmk almost totally abrogated the formation of apoptotic cells (Fig. 1b–d), corroborating that cell death represents genuine caspase-dependent apoptosis. Of note, treatment with Quer plus 2-DG resulted in free PI uptake by a high proportion of cells. Nevertheless, this effect was also suppressed by z-VAD-fmk (Fig. 1e), indicating that these cells represent late apoptosis (or “secondary necrosis”) instead of a genuine necrotic response. Finally, a time-course study (3–24 h) revealed that significant more than additive drug cooperation was firstly detectable at 6 h of treatment (approx. 15 % apoptosis in the combined treatment), and increased thereafter (see Additional file 1: Fig. S1). Thus, except when otherwise indicated, 6 h was the maximum time period used for further investigation of early regulatory events.
In a new set of experiments, 2-DG was combined with Cur (8 μM) and Gen (50 μM) instead of Quer. The suitability of these concentrations for combinatory studies in leukemia cell models was established in earlier publications [21, 29]. Some of the obtained results are presented in Fig. 2a, b. Cur alone caused negligible apoptosis but cooperated with 2-DG with similar efficacy as Quer (Fig. 2a). On the other hand, the efficacy of cooperation using 50 μM Gen was very low, and the concentration had to be increased to 100 μM (which is per se moderately lethal) to obtain a more satisfactory response (Fig. 2b). The results with both Cur and Gen were corroborated using annexin V/PI analysis (see Additional file 2: Fig. S2).
In addition, we examined the possible cooperation between Quer and 2-DG in NB4 promyelocytic and THP-1 promonocytic cells. Among other biochemical aspects, these cells exhibit marked differences in the metabolic phenotype: NB4 cells are highly dependent on glycolysis, while THP-1 cells may compensate glycolysis inhibition with fatty acid β-oxidation [30]. Accordingly, Fig. 2c, d shows that NB4 cells are more susceptible and THP-1 cells less susceptible to the cytotoxic action of 2-DG than HL60 cells. The figure also indicates more than additive cooperation between Quer and 2-DG in both cell lines, although with lower efficacy than in HL60 cells. Of note, the lower apoptotic efficacy was not due to a switch to a genuine necrotic response, since free PI uptake was almost totally suppressed by the caspase inhibitor (23.3 ± 1.1 vs. 5.8 ± 0.7 % positive cells upon treatment with 10 μM Quer plus 2.5 mM 2-DG in the absence and presence of z-VAD-fmk, respectively, in NB4 cells; and 19.3 ± 1.2 vs. 4.7 ± 1.0 % upon treatment with 20 μM Quer plus 20 mM 2-DG in the absence and the presence of z-VAD-fmk, respectively, in THP-1 cells).
It has been reported that flavonoids such as Quer and Gen may inhibit glucose and 2-DG uptake in different cell models, including AML cells, probably by affecting glucose transporter (GLUT-1) [31, 32]. For this reason new experiments were carried out using Lon, a HKII inhibitor structurally unrelated to 2-DG. The concentrations of 50 and 100 μM Lon were adopted as adequate for combinatory studies, according to our previous publications [11, 21]. The lethality of Lon alone was very low, but efficaciously cooperated with Quer, Cur and Gen to induce apoptosis in HL60 cells, as demonstrated by the frequency of cells with sub-G1 DNA content (Fig. 3a–c), and confirmed in the case of Quer plus Lon by the annexin V/PI assay (see Additional file 2: Fig. S2). The protective action of z-VAD-fmk corroborated again that cell death represented caspase dependent apoptosis (Fig. 3a, c), and that the concomitant free PI uptake was attributable to late apoptosis instead of genuine necrosis (38 ± 1.7 % vs 7.2 ± 0.7 PI-permeable cells upon treatment with 20 μM Quer plus 100 μM Lon in the absence and the presence of z-VAD-fmk, respectively).
Finally, we performed a set of experiments in which, instead of being treated with 2-DG, HL60 cells were cultured for 24 h in glucose–lacking (Glu−) medium, either in the absence or the presence of Quer. Both 2-DG treatment in complete (Glu+) medium and cell incubation in Glu medium resulted in partial depletion (approximately 40 % at 16 h) of intracellular ATP levels [12], and in cell proliferation inhibition (measured by cell counting), which in the case of Glu was exacerbated by addition of Quer (Fig. 3d). However, cells cultured in Glu were viable, and the generation of apoptosis by Quer/Glu was very low, in comparison to Quer plus 2-DG in complete medium (Fig. 3e). Again, the low apoptotic rate was not compensated by a switch to a necrotic response, as indicated by the low frequency of cells with free PI uptake (Fig. 3f).

Mitochondrial dysfunction

2-DG causes mitochondrial HKII inhibition and detachment and may therefore induce mIMP [7], as we previously corroborated in HL60 cells [12]. In addition, it has been reported that Quer may induce mIMP in isolated mitochondria by direct interaction with the adenine nucleotide translocase [33], although another study indicated both mIMP induction or inhibition, depending on the assay conditions [34]. In the present work, mitochondria dysfunction was firstly analyzed by measuring changes in ΔΨm after 14 h treatment with 20 μM Quer and 5 mM 2-DG, alone and in combination. The results are represented in Fig. 4a. Allowing for a slight ΔΨm increase by Quer alone, the most prominent effect was the appearance of a large subpopulation of cells with markedly low ΔΨm in the combined treatment. This subpopulation was suppressed by z-VAD-fmk, and hence likely represents the fraction of cells undergoing apoptosis. Then, a second set of experiments was carried out using the calcein/CoCl2 procedure, which is considered to provide a direct and more accurate determination of mIMP [35]. The results in Fig. 4b indicate that 2-DG (5 mM) and Quer (10–40 μM, in a concentration-dependent manner) caused mIMP induction, as evidenced by the decrease in calcein-derived fluorescence, and the decrease was augmented when both drugs were used in combination. Of note, this response was detected at 4 h of treatment, preceding the first manifestations of apoptosis execution (see Additional file 1: Fig. S1).

Oxidative stress

Although dietary polyphenols are normally considered as anti-oxidant, protective agents, there is ample evidence indicating that they may exert both anti-oxidant and pro-oxidant effects, depending on the chemical structure and treatment conditions. For instance, it was reported that Quer may either decrease [36] or increase [37, 38] ROS production in HL60 cells, and ROS increase mediated apoptosis induction [38]. For these reasons, we evaluated possible alterations in intracellular ROS accumulation after short treatments (3 h) with Quer and 2-DG, alone and in combination, using the ROS-sensitive fluorescent probe H2DCFDA. Some of the obtained results are presented in Fig. 5a. It was observed that 2-DG (5 mM) and Quer (10–40 μM, in a concentration-dependent manner) reduced the basal intracellular ROS content in HL60 cells, and the reduction was higher in the combined treatment. By contrast to Quer, Gen (50 μM) increased ROS levels (as previously reported [29]), but the increase was attenuated by combination with 2-DG. In a similar manner, and by contrast to 2-DG, Lon (100 μM) increased ROS content (as previously reported [11]), but this increase was totally abrogated by combination with Quer, reaching similar levels as with Quer alone. In summary, these results evidence large discrepancies in ROS production depending on the used treatment, which do not match and hence may not explain apoptosis potentiation in the combined treatments (see Figs. 13).
We previously reported that prolonged Quer treatment (14–24 h) reduced the intracellular GSH content in myeloid cells [20], while Gen and Cur were ineffective [21, 29]. In the present experiments we analyzed possible changes in GSH content upon short treatments of HL60 cells with Quer and the glycolytic inhibitors, using the GSH-sensitive fluorescent probe monochlorobimane. It was observed that GSH levels were not significantly affected by treatment for 3 h (data not shown) or 6 h (Fig. 5B) with 10–40 μM Quer, either alone or in combination with 2-DG (5 mM) or Lon (100 μM). Nonetheless, while these results exclude GSH as an early regulatory factor, we might not a priori exclude that a possible late depletion could have some effect on apoptosis progression. This possibility was indirectly investigated using BSO, a GSH specific synthesis inhibitor [39]. Treatment of HL60 cells with 1 mM BSO for 16 h caused a partial (35.3 ± 3.2 %) decrease in the basal GSH content, but the GSH inhibitor did not affect the proliferation rate (data not shown), nor caused cell lethality per se or in combination with 2-DG or Lon (Fig. 5c).

Protein kinase activation

As commented above (see ‘‘Background’’ section), we hypothesized that polyphenols might potentiate the apoptotic efficacy of glycolytic inhibitors by preventing the activation of defensive kinase pathways, either PI3K/Akt and/or MEK/ERK. To examine this hypothesis, in a first set of experiments we examined Akt and ERK phosphorylation/activation upon 1 and 6 h treatment of HL60 cells with Quer and 2-DG, alone and in combination (where, as previously indicated, Quer was applied 2 h before 2-DG). We also checked the response of S6-ribosomal protein (rpS6), which is downstream Akt; of GSK3α/β, which are phosphorylated by Akt and ERK [40, 41]; and of p38-MAPK, also described as a target of quercetin or quercetin-derived analogs in leukemia cells [42, 43]. The results, presented in Fig. 6, were as follows: (i) 2-DG (5 mM) stimulated Akt and rpS6 phosphorylation/activation, and the stimulation was abrogated or greatly attenuated by Quer. (ii) By contrast, ERK phosphorylation/activation was stimulated by both Quer and 2-DG, alone and in combination. (iii) Quer and 2-DG, alone and in combination, stimulated GSK3α/β phosphorylation (Ser21/9)/inactivation. Quer alone exerted higher effect on the α isoform, while 2-DG alone stimulated both isoforms. (iv) Quer, alone or with 2-DG, stimulated p38-MAPK phosphorylation/activation, while the effect of 2-DG alone was negligible.
Once we examined protein kinase modulation, the potential importance of these alterations was investigated using appropriate pharmacological inhibitors (the selected concentrations being adopted from our preceding studies with AML cells [44]). The results were as follows: (i) Co-treatment with the PI3K/Akt phosphorylation/activation inhibitors LY294002 (30 μM) or triciribine (10 μM) increased the apoptotic efficacy of 2-DG alone, and also augmented the slight apoptosis obtained with the combination of low concentrations of Quer (10 μM) plus 2-DG (2 mM) (Fig. 7a). This corroborates the role of Akt as a defensive kinase, and indicates that its inhibition by Quer may be at least in part responsible for the increased apoptosis in the combined (Quer/2-DG) treatment. (ii) Co-treatment with the MEK/ERK inhibitors PD98059 (20 μM) or U0126 (5 μM) increased the apoptotic efficacy of Quer and 2-DG alone, and also augmented apoptosis by the Quer/2-DG combination (Fig. 7b). This indicates that ERK functions as a defensive kinase serving to restrain lethality by Quer and 2-DG, but may not account for the increased apoptotic efficacy in the combined treatment. (iii) Co-treatment with the GSK-3 phosphorylation inhibitor SB216763 (10 μM) augmented the lethality of Quer and 2-DG alone and (although with lower efficacy) of the Quer plus 2-DG combination (Fig. 7c), indicating that drug-provoked GSK-3 phosphorylation/inactivation plays a defensive role. (iv) Finally, activation of p38-MAPK by Quer alone or Quer plus 2-DG seemed irrelevant for apoptosis, since the lethality was not modified by the pharmacologic inhibitors SB203580 (10 μM) (Fig. 7d) or BIRB 796 (0.1 μM: data not shown).
In final set of experiments, we analyzed Akt and ERK phosphorylation using other phenolic agents and glycolytic inhibitors, as previously assayed for apoptosis (see Figs. 2 and 3). The results are represented in Fig. 8a–c, and may be summarized as follows: (i) Gen (50 μM) produced qualitatively similar effects as Quer, namely rapid abrogation of 2-DG-provoked Akt phosphorylation, and (albeit with lower intensity) stimulation ERK activation when used alone. By contrast Cur (8 μM) was ineffective at 2 h, and required longer pre-treatment to suppress Akt activation. Moreover, under these conditions Cur, which did not per se affect ERK phosphorylation, prevented the activation of this kinase by 2-DG (Fig. 8a). (ii) Lon was assayed at 6 and 14 h, since we previously showed that kinase activations by this agent are delayed in relation to 2-DG [11]. It could be observed that Quer abrogated or greatly reduced Lon-provoked Akt activation at both time periods (Fig. 8b). (iii) Finally, by contrast to the strong stimulatory effect of 2-DG, the basal Akt phosphorylation was not affected by glucose starvation (Glu−), and co-incubation with Quer either did not affect (6 h) or caused a slight late decrease (20 h) (Fig. 8c). This correlates with apparent lack of lethality of glucose-starvation, and the low apoptotic efficacy of Quer under these conditions (Glu−/Quer) in comparison to 2-DG/Quer (see Fig. 3f). By contrast, ERK phosphorylation was strongly stimulated by the combined (Glu−/Quer) treatment (Fig. 8c).

Discussion

The present results indicate that that pre-treatment with low lethal concentrations of the flavonoid Quer strongly potentiates the anti-proliferative and apoptotic action of the glycolytic inhibitor 2-DG in HL60 AML cells. Apoptosis was assessed using different markers, namely DNA loss, phosphatidyl serine translocation, and ΔΨm decrease, and the protective action of z-VAD-fmk proves that it is in fact a caspase-mediated response. Our precedent studies demonstrated that Quer and 2-DG, as well as the other agents used here (Lon, Cur and Gen) activated apoptosis throughout the mitochondrial (“intrinsic”) executioner pathway [11, 12, 20, 21, 29]. In this study we show that 2-DG and Quer, both of them characterized as mitochondria-targeting drugs [7, 33], cause the rapid induction of mIMP (4 h), which may therefore represent a trigger or at least a necessary condition for apoptosis. Unfortunately the cause-effect relationship between mIMP and apoptosis could not be corroborated, due to the elevated toxicity of commonly used permeability transition pore modulators (cyclosporine A, bongkrekic acid) in leukemia cell models ([45], and our data not shown).
Additional experiments corroborated the cooperative effect using other polyphenols (Cur, Gen), anti-glycolytic agents (Lon), and leukemia cell models (NB4 promyelocytic, THP-1 promonocytic), although with different efficacy. For instance, Gen was less efficacious than Quer and Cur, and a moderately lethal concentration of the isoflavone was required to obtain good cooperation with 2-DG. As possible explanations, Gen causes cell arrest at G2/M (see Additional file 3: Fig. S3) and also stimulates myeloid cell differentiation [22], which might temporarily restrain the trigger of the apoptotic response. In the same manner, the efficacy of cooperation between Quer and 2-DG was lower in NB4 and THP1 than in HL60 cells. This might be explained by intrinsic differences in molecular and biochemical properties and in the maturation stage of these cell lines. For instance, in addition to the above indicated different susceptibility to 2-DG, THP-1 promonocytic cells are more resistant to Quer [46] (an in our experience also to other cytotoxic agents) than the less mature HL60 and NB4 cells. On the other hand, the finding that Quer plus 2-DG and Quer plus Lon induced apoptosis with similar efficacy allows to exclude that apoptosis potentiation may be a trivial consequence of polyphenol-provoked inhibition of GLUT 1 activity and 2-DG uptake, and hence of subsequent biochemical responses (e.g., 2-DG-provoked activation of defensive kinases, as discussed later). By contrast, the apoptotic efficacy of Quer was only marginally increased when 2-DG treatment was substituted by glucose starvation. This indicates that the partial ATP depletion provoked by both treatments may affect cell proliferation, but it is not a determinant of drug lethality. Finally, a cautionary note must be expressed. We centered the attention on the apoptotic response, and for convenience (better observation of drug cooperation) selected sub-lethal drug concentrations. Nonetheless these concentrations caused appreciable anti-proliferative effects, as measured by the MTT assay (Fig. 1a) or by cell counting (the proliferation inhibition rates at 24 h, in relation to the control, were: 43 % for 5 mM 2-DG; 5 % by 100 μM Lon; 44 % by 20 μM Quer; 21 % by 8 μM Cur; and 65 % by 50 μM Gen). This may be explained in some cases by cell cycle disruption (e.g., almost total G2/M arrest by Gen, and with lower intensity by Quer: see Additional file 3: Fig. S3). The molecular mechanisms responsible for Gen-provoked cycle arrest were investigated in a preceding article [22]. In addition, a possible activation of autophagy, which opposes apoptosis, may not be discarded.
The generation of apoptosis and/or necrosis by anti-tumour drugs is frequently associated to oxidative stress, two manifestations of which are the increase in intracellular ROS accumulation and the decrease in anti-oxidant molecules such as GSH. For instance, we earlier demonstrated that the potentiation of arsenic trioxide (Trisenox)-provoked apoptosis by Lon and Gen was mediated by the stimulation of ROS production [11, 29], probably due to the interference of Lon and Gen with the mitochondrial respiratory chain [47, 48]. On the other hand, 2-DG was reported to decrease ROS [49, 50], while Quer may exert either inhibitory [33] stimulatory [37, 38] effects. Our present results confirm the drug-dependent variability, and above all prove that the increase in lethality in the combined treatments (Quer plus 2-DG and Quer plus Lon) may not be adequately explained by ROS over-production. Thus, Quer not only reduced the basal ROS content, but exacerbated the decrease caused by 2-DG, and reversed the increase caused by Lon to levels lower than in untreated cells. In another study we observed that prolonged (14–24 h) treatment with Quer decreased the intracellular GSH content, and as a consequence potentiated the lethality of the GSH-sensitive drug arsenic trioxide [20]. However a regulatory role of GSH may be excluded in the present conditions. In fact, Quer did not cause early changes in GSH content, and the impact on apoptosis of a potential long-term alteration may be discarded, in view of the lack of effects of GSH specific inhibitor BSO.
Finally, the present results corroborate the capacity of 2-DG and Lon to stimulate Akt activation in AML cells, and demonstrate that the stimulation is abrogated or attenuated by co-treatment with Quer, Gen and Cur, which at the same time potentiate apoptosis. The cause-effect relationship between Akt inhibition and apoptosis potentiation was supported by experiments using standard pharmacologic protein kinase inhibitors, and was also strengthened by the results obtained using glucose-free medium, where the lack of Akt activation correlates with poor Quer lethality. There are nevertheless some quantitative differences: thus, Akt activation by Lon was delayed in relation to 2-DG; a prolonged pre-treatment with Cur (14 h) was required to abrogate 2-DG-provoked Akt activation; and 50 μM Gen sufficed to block 2-DG-provoked Akt activation, although the pro-apoptotic action of this Gen concentration was poor. The discrepancies might be explained by the particular action mechanisms of the used drugs, and also indicate that Akt inhibition is an important factor but not the only one accounting for apoptotic potentiation in the combined treatments. Noteworthy, prolonged pre-treatment with Cur also abrogated 2-DG-provoked ERK activation, indicating that at least in some circumstances this kinase also regulates apoptosis potentiation in the combined treatments. Nonetheless the function of ERKs in the present experiments is less clear, mainly because the disparity of effects caused by polyphenols. Thus, ERK phosphorylation was strongly stimulated by Quer, slightly stimulated by Gen, and unaffected by Cur. Of note, co-treatment with MEK/ERK inhibitors increased the apoptotic efficacy of Quer alone. It seems therefore that ERK activation serves to restrain excessive polyphenol toxicity, in the same manner as 2-DG toxicity. In addition Quer, alone or with 2-DG, caused p38-MAP activation, but this response seems irrelevant for apoptosis as judged by the null effect of kinase inhibitors. This conclusion is consistent with results obtained by other authors using a Quer analogue [43]. Finally, earlier reports indicated that 2-DG elicits Akt-dependent GSK-3β phosphorylation [13] while quercetin increases [51] or does not affect [52] kinase phosphorylation. Using pharmacologic inhibitors, other articles proved that GSK-3 regulates cell growth and/or apoptosis in leukemia cells [5355]. Our present results indicate that Quer and 2-DG cause hyper-phosphorylation (Ser21/9)/inactivation of GSK-3α/β in HL60 cells, although with certain drug-specificity. Thus, Quer (which only activated ERKs) preferentially stimulated the α isoform, while 2-DG (which activated Akt and ERKs) stimulated both α and β isoforms. Most studies in the literature only centered the attention on GSK-3β, but some reports call attention on the functional importance of the α isoform. As an example, GSK-3α knock-down reduced proliferation and caused spontaneous apoptosis [56], and potentiated bortezomib-induced toxicity [57], in leukemia cells. Our results show that the GSK-3 inhibitor SB216763 (10 μM) does no cause per se significant lethality, but potentiates apoptosis induction by Quer and 2-DG, alone and in combination, in HL60 cells. Since as commented above Quer preferentially stimulates GSK-3α, we might postulate that this isoform is the most important for regulation of Quer lethality. Unfortunately the lack of specificity of the up to date available pharmacologic inhibitors, and the difficulty to perform satisfactory knock-down procedures in the used leukemia cell model, impeded us until now to obtain more clear conclusions.

Conclusions

De-regulation of the MEK/ERK and especially the PI3K/Akt pathway are among the most frequent alterations associated to the generation of the tumour phenotype, as well as to the intrinsic or acquired resistance of tumour cells to anti-cancer treatments. Dietary phenolic agents are in general well tolerated, and might therefore be preferable to approach the problem instead of synthetic pharmacologic drugs. The present results indicate that co-treatment with low concentrations of selected polyphenols increase the apoptotic efficacy of the glycolytic inhibitors 2-DG and Lon in human acute AML cell models, and that this effect may be at least in part explained by the prevention of defensive protein kinase activation, mainly Akt and in some circumstances ERKs. Of course, we did not attempt performing an exhaustive study: in fact, polyphenols affect multiple biochemical and molecular mechanisms other than Akt inhibition [1618], 2-DG is a dual glycolysis and N-glycosylation inhibitor [4], and Lon inhibits glycolysis but also lactate transport (causing intracellular acidification) and mitochondrial respiration [5, 48]. In spite of the conceptual and technical limitations we believe that our in vitro study may offer some ideas to improve the efficacy of this potentially important group of anti-tumour drugs.

Authors’ contributions

All authors performed experiments. EB and PA analyzed the data. PA wrote the manuscript. EB, MCE, MCGF, JR and MCBA discussed and corrected the manuscript. All authors read and approved the final manuscript.

Acknowledgements

The authors acknowledge Profs. J. León and M.D. Delgado (Departamento de Biología Molecular, Facultad de Medicina, Universidad de Cantabria, Santander, Spain) for generous gift of NB4 cells, and to Dr. Pedro Lastres and Mr. Guillermo Padilla (Servicio de Citometría de Flujo and Servicio de Bioinformática y Bioestadística, respectively, Centro de Investigaciones Biológicas, CSIC, Madrid) for help with flow cytometry and statistical analysis.

Competing interests

The authors declare that they have no competing interests.

Availability of data and materials

Not applicable.
Not applicable.
Not applicable.

Funding

This work was supported by grant SAF2010-20256 (to P.A.) from the Spanish Ministerio de Ciencia e Innovación, Plan Nacional de Investigación Científica, Desarrollo e Innovación Tecnológica, Dirección General de Investigación (http://​www.​mineco.​gob.​es/​portal/​site/​mineco/​idi). M.C.E. was recipient of a JAE-Predoc fellowship from the Consejo Superior de Investigaciones Científicas (http://​www.​csic.​es/​web/​guest/​programa-jae). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.
Open AccessThis article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://​creativecommons.​org/​licenses/​by/​4.​0/​), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://​creativecommons.​org/​publicdomain/​zero/​1.​0/​) applies to the data made available in this article, unless otherwise stated.
Literatur
1.
Zurück zum Zitat Kroemer G, Pouyssegur J. Tumor cell metabolism: cancer’s Achilles’ heel. Cancer Cell. 2008;13:472–82.CrossRefPubMed Kroemer G, Pouyssegur J. Tumor cell metabolism: cancer’s Achilles’ heel. Cancer Cell. 2008;13:472–82.CrossRefPubMed
2.
Zurück zum Zitat Tennant DA, Duran RV, Gottlieb E. Targeting metabolic transformation for cancer therapy. Nat Rev Cancer. 2010;10:267–77.CrossRefPubMed Tennant DA, Duran RV, Gottlieb E. Targeting metabolic transformation for cancer therapy. Nat Rev Cancer. 2010;10:267–77.CrossRefPubMed
4.
Zurück zum Zitat Zhang D, Li J, Wanh F, Hu J, Wang S, Sun Y. 2-Deoxy-d-glucose targeting of glucose metabolism in cancer cells as a potential therapy. Cancer Let. 2014;355:176–83.CrossRef Zhang D, Li J, Wanh F, Hu J, Wang S, Sun Y. 2-Deoxy-d-glucose targeting of glucose metabolism in cancer cells as a potential therapy. Cancer Let. 2014;355:176–83.CrossRef
5.
Zurück zum Zitat Di Cosimo S, Ferretti G, Papaldo P, Carlini P, Fabi A, Cognetti F. Lonidamine: efficacy and safety in clinical trials for the treatment of solid tumors. Drugs Today. 2003;39:157–74.CrossRefPubMed Di Cosimo S, Ferretti G, Papaldo P, Carlini P, Fabi A, Cognetti F. Lonidamine: efficacy and safety in clinical trials for the treatment of solid tumors. Drugs Today. 2003;39:157–74.CrossRefPubMed
6.
Zurück zum Zitat Ganaphaty-Kanniappan S, Kunjithapatham R, Geswchwind JF. Anticancer efficacy of the metabolic blocker 3-bromopyruvate: specific molecular targeting. Anticancer Res. 2013;33:13–20. Ganaphaty-Kanniappan S, Kunjithapatham R, Geswchwind JF. Anticancer efficacy of the metabolic blocker 3-bromopyruvate: specific molecular targeting. Anticancer Res. 2013;33:13–20.
7.
Zurück zum Zitat Chiara F, Castellaro D, Marin O, Petronilli V, Brusilow WS, Juhaszova M, et al. Hexokinase II detachment from mitochondria triggers apoptosis through the permeability transition pore independent of voltage-dependent anion channels. PLoS One. 2008;3:1–11.CrossRef Chiara F, Castellaro D, Marin O, Petronilli V, Brusilow WS, Juhaszova M, et al. Hexokinase II detachment from mitochondria triggers apoptosis through the permeability transition pore independent of voltage-dependent anion channels. PLoS One. 2008;3:1–11.CrossRef
8.
Zurück zum Zitat Pereira da Silva AP, El-Bacha T, Kyaw N, dos Santos RS, da-Silva WS, Almeida FC, et al. Inhibition of energy-producing pathways of HepG2 cells by 3-bromopyruvate. Biochem J. 2009;417:717–26.CrossRefPubMed Pereira da Silva AP, El-Bacha T, Kyaw N, dos Santos RS, da-Silva WS, Almeida FC, et al. Inhibition of energy-producing pathways of HepG2 cells by 3-bromopyruvate. Biochem J. 2009;417:717–26.CrossRefPubMed
9.
Zurück zum Zitat Dwarakanath BS, Jain V. Targeting glucose metabolism with 2-deoxy-d-glucose for improving cancer therapy. Future Oncol. 2009;5:581–5.CrossRefPubMed Dwarakanath BS, Jain V. Targeting glucose metabolism with 2-deoxy-d-glucose for improving cancer therapy. Future Oncol. 2009;5:581–5.CrossRefPubMed
10.
Zurück zum Zitat Ki YH, Verhoeven HA, Lee MJ, Corbin DJ, Vogl TJ, Pedersen PL. A translational study “case report” on the small molecule “energy blocker” 3-bromopyruvate (3BP) as a potent anticancer agent: from bench side to bedside. J Bioenerg Biomembr. 2012;44:163–70.CrossRef Ki YH, Verhoeven HA, Lee MJ, Corbin DJ, Vogl TJ, Pedersen PL. A translational study “case report” on the small molecule “energy blocker” 3-bromopyruvate (3BP) as a potent anticancer agent: from bench side to bedside. J Bioenerg Biomembr. 2012;44:163–70.CrossRef
11.
Zurück zum Zitat Calviño E, Estañ MC, Simon GP, Sancho P, Boyano-Adanez MC, de Blas E, et al. Increased apoptotic efficacy of lonidamine plus arsenic trioxide combination in human leukemia cells. Reactive oxygen species generation and defensive protein kinase (MEK/ERK, Akt/mTOR) modulation. Biochem Pharmacol. 2011;82:1619–29.CrossRefPubMed Calviño E, Estañ MC, Simon GP, Sancho P, Boyano-Adanez MC, de Blas E, et al. Increased apoptotic efficacy of lonidamine plus arsenic trioxide combination in human leukemia cells. Reactive oxygen species generation and defensive protein kinase (MEK/ERK, Akt/mTOR) modulation. Biochem Pharmacol. 2011;82:1619–29.CrossRefPubMed
12.
Zurück zum Zitat Estañ MC, Calviño E, de Blas E, Boyano-Adanez MC, Mena ML, Gómez-Gómez M, et al. 2-Deoxy-d-glucose cooperates with arsenic trioxide to induce apoptosis in leukemia cells: involvement of IGF-1R-regulated Akt/mTOR, MEK/ERK and LKB-1/AMPK signaling pathways. Biochem Pharmacol. 2012;84:1604–16.CrossRefPubMed Estañ MC, Calviño E, de Blas E, Boyano-Adanez MC, Mena ML, Gómez-Gómez M, et al. 2-Deoxy-d-glucose cooperates with arsenic trioxide to induce apoptosis in leukemia cells: involvement of IGF-1R-regulated Akt/mTOR, MEK/ERK and LKB-1/AMPK signaling pathways. Biochem Pharmacol. 2012;84:1604–16.CrossRefPubMed
13.
Zurück zum Zitat Zhong D, Liu X, Schafer-Hales K, Marcus AI, Khuri FR, Sun SY, et al. 2-Deoxyglucose induces Akt phosphorylation via a mechanism independent of LKB1/AMP-activated protein kinase signaling activation or glycolysis inhibition. Mol Cancer Ther. 2008;7:809–17.CrossRefPubMed Zhong D, Liu X, Schafer-Hales K, Marcus AI, Khuri FR, Sun SY, et al. 2-Deoxyglucose induces Akt phosphorylation via a mechanism independent of LKB1/AMP-activated protein kinase signaling activation or glycolysis inhibition. Mol Cancer Ther. 2008;7:809–17.CrossRefPubMed
14.
Zurück zum Zitat Zhong D, Xiong L, Liu T, Liu X, Liu X, Chen J, et al. The glycolytic inhibitor 2-deoxyglucose activates multiple prosurvival pathways through IGF1R. J Biol Chem. 2009;284:23225–33.CrossRefPubMedPubMedCentral Zhong D, Xiong L, Liu T, Liu X, Liu X, Chen J, et al. The glycolytic inhibitor 2-deoxyglucose activates multiple prosurvival pathways through IGF1R. J Biol Chem. 2009;284:23225–33.CrossRefPubMedPubMedCentral
15.
Zurück zum Zitat De Salvo J, Kuznestsov JN, Du J, Leclerk GM, Lampidis TJ, Barredo JC. Inhibition of Akt potentiates 2-DG-induced apoptosis via downregulation of UPR in acute lymphoblastic leukemia. Mol Cancer Res. 2012;10:969–78.CrossRef De Salvo J, Kuznestsov JN, Du J, Leclerk GM, Lampidis TJ, Barredo JC. Inhibition of Akt potentiates 2-DG-induced apoptosis via downregulation of UPR in acute lymphoblastic leukemia. Mol Cancer Res. 2012;10:969–78.CrossRef
16.
Zurück zum Zitat Fantini M, Benvenuto M, Masuelli L, Frajese GV, Tresoldi I, Modesti A, et al. In vitro and in vivo antitumoural effects of combinations of polyphenols or polyphenols and anticancer drugs: perspectives on cancer. Int J Mol Sci. 2015;16:9236–82.CrossRefPubMedPubMedCentral Fantini M, Benvenuto M, Masuelli L, Frajese GV, Tresoldi I, Modesti A, et al. In vitro and in vivo antitumoural effects of combinations of polyphenols or polyphenols and anticancer drugs: perspectives on cancer. Int J Mol Sci. 2015;16:9236–82.CrossRefPubMedPubMedCentral
17.
Zurück zum Zitat Ramos S. Cancer chemoprevention and chemotherapy: dietary polyphenols and signalling pathways. Mol Nutr Food Res. 2008;52:507–26.CrossRefPubMed Ramos S. Cancer chemoprevention and chemotherapy: dietary polyphenols and signalling pathways. Mol Nutr Food Res. 2008;52:507–26.CrossRefPubMed
18.
Zurück zum Zitat Prasad S, Phromnoi K, Yadav VR, Chaturvedi MM, Aggarwal BB. Targeting inflammatory pathways by flavonoids for prevention and treatment of cancer. Planta Med. 2010;76:1044–63.CrossRefPubMed Prasad S, Phromnoi K, Yadav VR, Chaturvedi MM, Aggarwal BB. Targeting inflammatory pathways by flavonoids for prevention and treatment of cancer. Planta Med. 2010;76:1044–63.CrossRefPubMed
20.
Zurück zum Zitat Ramos AM, Aller P. Quercetin decreases intracellular GSH content and potentiates the apoptotic action of the antileukemic drug arsenic trioxide in human leukemia cell lines. Biochem Pharmacol. 2008;75:1912–23.CrossRefPubMed Ramos AM, Aller P. Quercetin decreases intracellular GSH content and potentiates the apoptotic action of the antileukemic drug arsenic trioxide in human leukemia cell lines. Biochem Pharmacol. 2008;75:1912–23.CrossRefPubMed
21.
Zurück zum Zitat Sánchez Y, Simón GP, Calviño E, de Blas E, Aller P. Curcumin stimulates reactive oxygen species production and potentiates apoptosis induction by the antitumour drugs arsenic trioxide and lonidamine in human myeloid leukemia cell lines. J Pharmacol Exp Ther. 2010;335:114–23.CrossRefPubMed Sánchez Y, Simón GP, Calviño E, de Blas E, Aller P. Curcumin stimulates reactive oxygen species production and potentiates apoptosis induction by the antitumour drugs arsenic trioxide and lonidamine in human myeloid leukemia cell lines. J Pharmacol Exp Ther. 2010;335:114–23.CrossRefPubMed
22.
Zurück zum Zitat Sánchez Y, Amrán D, de Blas E, Aller P. Regulation of genistein-induced differentiation in human myeloid leukaemia cells (HL60, NB4). Protein kinase modulation and reactive oxygen species generation. Biochem Pharmacol. 2009;77:384–96.CrossRefPubMed Sánchez Y, Amrán D, de Blas E, Aller P. Regulation of genistein-induced differentiation in human myeloid leukaemia cells (HL60, NB4). Protein kinase modulation and reactive oxygen species generation. Biochem Pharmacol. 2009;77:384–96.CrossRefPubMed
23.
Zurück zum Zitat Collins SJ, Gallo RC, Gallagher RE. Continuous growth and differentiation of human myeloid leukaemic cells in suspension culture. Nature. 1977;270:347–9.CrossRefPubMed Collins SJ, Gallo RC, Gallagher RE. Continuous growth and differentiation of human myeloid leukaemic cells in suspension culture. Nature. 1977;270:347–9.CrossRefPubMed
24.
Zurück zum Zitat Tsuchiya S, Yamabe M, Yamaguchi Y, Kobayashi Y, Konno T, Tada K. Establishment and characterization of a human acute monocytic leukemia cell line (THP-1). Int J Cancer. 1980;26:171–6.CrossRefPubMed Tsuchiya S, Yamabe M, Yamaguchi Y, Kobayashi Y, Konno T, Tada K. Establishment and characterization of a human acute monocytic leukemia cell line (THP-1). Int J Cancer. 1980;26:171–6.CrossRefPubMed
25.
Zurück zum Zitat Lanotte M, Martin-Thouvenin V, Najman S, Balerini P, Valensi F, Berger R. NB4, a maturation inducible cell line with t(15;17) marker isolated from a human acute promyelocytic leukemia (M3). Blood. 1991;77:1080–6.PubMed Lanotte M, Martin-Thouvenin V, Najman S, Balerini P, Valensi F, Berger R. NB4, a maturation inducible cell line with t(15;17) marker isolated from a human acute promyelocytic leukemia (M3). Blood. 1991;77:1080–6.PubMed
26.
Zurück zum Zitat Troyano A, Fernández C, Sancho P, de Blas E, Aller P. Effect of glutathione depletion on antitumor drug toxicity (apoptosis and necrosis) in U-937 human promonocytic cells. The role of intracellular oxidation. J Biol Chem. 2001;50:47107–15.CrossRef Troyano A, Fernández C, Sancho P, de Blas E, Aller P. Effect of glutathione depletion on antitumor drug toxicity (apoptosis and necrosis) in U-937 human promonocytic cells. The role of intracellular oxidation. J Biol Chem. 2001;50:47107–15.CrossRef
27.
Zurück zum Zitat Petronilli V, Miotto G, Canton M, Brini M, Colonna R, Bernardi P, et al. Transient and long-lasting openings of the mitochondrial permeability transition pore can be monitored directly in intact cells by changes in mitochondrial calcein fluorescence. Biophys J. 1999;76:725–34.CrossRefPubMedPubMedCentral Petronilli V, Miotto G, Canton M, Brini M, Colonna R, Bernardi P, et al. Transient and long-lasting openings of the mitochondrial permeability transition pore can be monitored directly in intact cells by changes in mitochondrial calcein fluorescence. Biophys J. 1999;76:725–34.CrossRefPubMedPubMedCentral
28.
Zurück zum Zitat Galan A, Garcia-Bermejo ML, Troyano A, Vilaboa NE, de Blas E, Kazanietz MG, et al. Stimulation of p38 mitogen-activated protein kinase is an early regulatory event for the cadmium-induced apoptosis in human promonocytic cells. J Biol Chem. 2000;275:11418–24.CrossRefPubMed Galan A, Garcia-Bermejo ML, Troyano A, Vilaboa NE, de Blas E, Kazanietz MG, et al. Stimulation of p38 mitogen-activated protein kinase is an early regulatory event for the cadmium-induced apoptosis in human promonocytic cells. J Biol Chem. 2000;275:11418–24.CrossRefPubMed
29.
Zurück zum Zitat Sánchez Y, Amrán D, Fernández C, de Blas E, Aller P. Genistein selectively potentiates arsenic trioxide-induced apoptosis in human leukemia cells via reactive oxygen species generation and activation of reactive oxygen species-inducible protein kinases (p38-MAPK, AMPK). Int J Cancer. 2008;123:1205–14.CrossRefPubMed Sánchez Y, Amrán D, Fernández C, de Blas E, Aller P. Genistein selectively potentiates arsenic trioxide-induced apoptosis in human leukemia cells via reactive oxygen species generation and activation of reactive oxygen species-inducible protein kinases (p38-MAPK, AMPK). Int J Cancer. 2008;123:1205–14.CrossRefPubMed
30.
Zurück zum Zitat Suganuma K, Miwa H, Imai N, Shikami M, Goto M, Mizuno S, et al. Energy metabolism of leukemia cells: glycolysis versus oxidative phosphorylation. Leuk Lymphoma. 2010;51:2112–9.CrossRefPubMed Suganuma K, Miwa H, Imai N, Shikami M, Goto M, Mizuno S, et al. Energy metabolism of leukemia cells: glycolysis versus oxidative phosphorylation. Leuk Lymphoma. 2010;51:2112–9.CrossRefPubMed
31.
Zurück zum Zitat Park JB. Flavonoids are potential inhibitors of glucose uptake in U937 cells. Biochem Biophys Res Commun. 1999;260:568–74.CrossRefPubMed Park JB. Flavonoids are potential inhibitors of glucose uptake in U937 cells. Biochem Biophys Res Commun. 1999;260:568–74.CrossRefPubMed
32.
Zurück zum Zitat Vera JC, Reyes AM, Velásques FV, Rivas CI, Zhang RH, Strobel P, et al. Direct inhibition of the hexose transporter GLUT1 by tyrosine kinase inhibitors. Biochemistry. 2001;40:777–90.CrossRefPubMed Vera JC, Reyes AM, Velásques FV, Rivas CI, Zhang RH, Strobel P, et al. Direct inhibition of the hexose transporter GLUT1 by tyrosine kinase inhibitors. Biochemistry. 2001;40:777–90.CrossRefPubMed
33.
Zurück zum Zitat Ortega R, García N. The flavonoid quercetin induces changes in mitochondrial permeability by inhibiting adenine nucleotide translocase. J Bioenerg Biomembr. 2009;41:41–7.CrossRefPubMed Ortega R, García N. The flavonoid quercetin induces changes in mitochondrial permeability by inhibiting adenine nucleotide translocase. J Bioenerg Biomembr. 2009;41:41–7.CrossRefPubMed
34.
Zurück zum Zitat De Marchi U, Biasutto L, Garbisa S, Toninello A, Zoratti M. Quercetin can act either as an inhibitor or an inducer of the mitochondrial permeability transition pore: a demonstration of the ambivalent redox character of polyphenols. Biochim Biophys Acta. 2009;1787:1425–32.CrossRefPubMed De Marchi U, Biasutto L, Garbisa S, Toninello A, Zoratti M. Quercetin can act either as an inhibitor or an inducer of the mitochondrial permeability transition pore: a demonstration of the ambivalent redox character of polyphenols. Biochim Biophys Acta. 2009;1787:1425–32.CrossRefPubMed
35.
Zurück zum Zitat Poncet D, Boya P, Métivier D, Zamzami N, Kroemer G. Cytofluorometric quantitation of apoptosis-driven inner mitochondrial membrane permeabilization. Apoptosis. 2003;8:521–30.CrossRefPubMed Poncet D, Boya P, Métivier D, Zamzami N, Kroemer G. Cytofluorometric quantitation of apoptosis-driven inner mitochondrial membrane permeabilization. Apoptosis. 2003;8:521–30.CrossRefPubMed
36.
Zurück zum Zitat Bestwick CS, Milne L. Quercetin modifies reactive oxygen levels but exerts only partial protection against oxidative stress within HL-60 cells. Biochim Biophys Acta. 2001;1528:49–59.CrossRefPubMed Bestwick CS, Milne L. Quercetin modifies reactive oxygen levels but exerts only partial protection against oxidative stress within HL-60 cells. Biochim Biophys Acta. 2001;1528:49–59.CrossRefPubMed
37.
Zurück zum Zitat Wang IK, Lin-Shiau SY, Lin JK. Induction of apoptosis by apigenin and related flavonoids through cytochorme c release and activation of caspase-9 and caspase-3 in leukaemia HL-60 cells. Eur J Cancer. 1999;35:1517–25.CrossRefPubMed Wang IK, Lin-Shiau SY, Lin JK. Induction of apoptosis by apigenin and related flavonoids through cytochorme c release and activation of caspase-9 and caspase-3 in leukaemia HL-60 cells. Eur J Cancer. 1999;35:1517–25.CrossRefPubMed
38.
Zurück zum Zitat Lee WJ, Hsiao M, Chang JL, Yang SF, Tseng TH, Cheng CW, et al. Quercetin induces mitochondrial-derived apoptosis via reactive oxygen species-mediated ERK activation in HL-60 leukemia cells and xenograft. Arch Toxicol. 2015;89:1103–17.CrossRefPubMed Lee WJ, Hsiao M, Chang JL, Yang SF, Tseng TH, Cheng CW, et al. Quercetin induces mitochondrial-derived apoptosis via reactive oxygen species-mediated ERK activation in HL-60 leukemia cells and xenograft. Arch Toxicol. 2015;89:1103–17.CrossRefPubMed
39.
Zurück zum Zitat Bailey HH. L-S, R-buthionine sulfoximine: historical development and clinical issues. Chem Biol Interact. 1998;111–112:239–54.CrossRefPubMed Bailey HH. L-S, R-buthionine sulfoximine: historical development and clinical issues. Chem Biol Interact. 1998;111–112:239–54.CrossRefPubMed
41.
Zurück zum Zitat Maurer U, Preiss F, Brauns-Schubert P, Schlicher L, Charvert C. GSK-3—at the crossroads of cell death and survival. J Cell Sci. 2014;127:1369–78.CrossRefPubMed Maurer U, Preiss F, Brauns-Schubert P, Schlicher L, Charvert C. GSK-3—at the crossroads of cell death and survival. J Cell Sci. 2014;127:1369–78.CrossRefPubMed
42.
Zurück zum Zitat Mertens-Talkot SU, Bomser JA, Romero C, Talcott ST, Percival SS. Ellagic acid potentiates the effect of quercetin on p21waf1/cip1, p53, and MAP-kinases without affecting intracellular generation of reactive oxygen species in vitro. J Nutr. 2005;135:609–14. Mertens-Talkot SU, Bomser JA, Romero C, Talcott ST, Percival SS. Ellagic acid potentiates the effect of quercetin on p21waf1/cip1, p53, and MAP-kinases without affecting intracellular generation of reactive oxygen species in vitro. J Nutr. 2005;135:609–14.
43.
Zurück zum Zitat Rubio S, Quintana J, Eiroa JL, Triana J, Estévez F. Acetyl derivative of quercetin 3-methyl ether-induced cell death in human leukemia cells is amplified by the inhibition of ERK. Carcinogenesis. 2007;28:2105–13.CrossRefPubMed Rubio S, Quintana J, Eiroa JL, Triana J, Estévez F. Acetyl derivative of quercetin 3-methyl ether-induced cell death in human leukemia cells is amplified by the inhibition of ERK. Carcinogenesis. 2007;28:2105–13.CrossRefPubMed
44.
Zurück zum Zitat Calviño E, Estañ MC, Sanchez-Martin C, Brea R, de Blas E, MdelC Boyano-Adánez, et al. Regulation of death induction and chemosensitizing action of 3-bromopyruvate in myeloid leukemia cells: energy depletion, oxidative stress, and protein kinase activity modulation. J Pharmacol Exp Ther. 2014;348:324–35.CrossRefPubMed Calviño E, Estañ MC, Sanchez-Martin C, Brea R, de Blas E, MdelC Boyano-Adánez, et al. Regulation of death induction and chemosensitizing action of 3-bromopyruvate in myeloid leukemia cells: energy depletion, oxidative stress, and protein kinase activity modulation. J Pharmacol Exp Ther. 2014;348:324–35.CrossRefPubMed
45.
Zurück zum Zitat Zunino SJ, Storms DH. Resveratrol-induced apoptosis is enhanced in acute lymphoblastic leukemia cells by modulation of the mitochondrial permeability transition pore. Cancer Lett. 2006;240:123–34.CrossRefPubMed Zunino SJ, Storms DH. Resveratrol-induced apoptosis is enhanced in acute lymphoblastic leukemia cells by modulation of the mitochondrial permeability transition pore. Cancer Lett. 2006;240:123–34.CrossRefPubMed
46.
Zurück zum Zitat Shen SC, Chen YC, Hsu FL, Lee WR. Differential apoptosis-inducing effect of quercetin and its glycosides in human promyeloleukemic HL-60 cells by alternative activation of the caspase 3 cascade. J Cell Biochem. 2003;89:1044–55.CrossRefPubMed Shen SC, Chen YC, Hsu FL, Lee WR. Differential apoptosis-inducing effect of quercetin and its glycosides in human promyeloleukemic HL-60 cells by alternative activation of the caspase 3 cascade. J Cell Biochem. 2003;89:1044–55.CrossRefPubMed
47.
Zurück zum Zitat Salvi M, Brunati AM, Clari G, Toninello A. Interaction of genistein with the mitochondrial electron transport chain results in opening of the membrane transition pore. Biochim Biophys Acta. 2002;1556:187–96.CrossRefPubMed Salvi M, Brunati AM, Clari G, Toninello A. Interaction of genistein with the mitochondrial electron transport chain results in opening of the membrane transition pore. Biochim Biophys Acta. 2002;1556:187–96.CrossRefPubMed
48.
Zurück zum Zitat Guo L, Shestov AA, Worth AJ, Nath K, Nelson DS, Leeper DB, et al. Inhibition of mitochondrial complex II by the anticancer agent lonidamine. J Biol Chem. 2016;291:42–57.CrossRefPubMed Guo L, Shestov AA, Worth AJ, Nath K, Nelson DS, Leeper DB, et al. Inhibition of mitochondrial complex II by the anticancer agent lonidamine. J Biol Chem. 2016;291:42–57.CrossRefPubMed
49.
Zurück zum Zitat Duan W, Mattson MP. Dietary restriction and 2-deoxyglucose administration improve behavioral outcome and reduce degeneration of dopaminergic neurons in models of Parkinson’s disease. J Neurosci Res. 1999;57:195–206.CrossRefPubMed Duan W, Mattson MP. Dietary restriction and 2-deoxyglucose administration improve behavioral outcome and reduce degeneration of dopaminergic neurons in models of Parkinson’s disease. J Neurosci Res. 1999;57:195–206.CrossRefPubMed
50.
Zurück zum Zitat Xi H, Barredo JC, Merchan JR, Lampidis TJ. Endoplasmic reticulum stress induced by 2-deoxyglucose but not glucose starvation activates AMPK through CaMKKβ leading to autophagy. Biochem Pharmacol. 2013;85:1463–77.CrossRefPubMed Xi H, Barredo JC, Merchan JR, Lampidis TJ. Endoplasmic reticulum stress induced by 2-deoxyglucose but not glucose starvation activates AMPK through CaMKKβ leading to autophagy. Biochem Pharmacol. 2013;85:1463–77.CrossRefPubMed
51.
Zurück zum Zitat Lee KH, Yoo CG. Simultaneous inactivation of GSK-3β suppresses quercetin-induced apoptosis by inhibiting the JNK pathway. Am J Physiol Lung Cell Mol Physiol. 2013;304:L782–9.CrossRefPubMed Lee KH, Yoo CG. Simultaneous inactivation of GSK-3β suppresses quercetin-induced apoptosis by inhibiting the JNK pathway. Am J Physiol Lung Cell Mol Physiol. 2013;304:L782–9.CrossRefPubMed
52.
Zurück zum Zitat Pahlke G, Ngiewith Y, Kern M, Jakobs S, Marko D, Eisenbrand G. Impact of quercetin and EGCG on key elements if the Wnt pathway in human colon carcinoma cells. J Agric Food Chem. 2006;54:7075–82.CrossRefPubMed Pahlke G, Ngiewith Y, Kern M, Jakobs S, Marko D, Eisenbrand G. Impact of quercetin and EGCG on key elements if the Wnt pathway in human colon carcinoma cells. J Agric Food Chem. 2006;54:7075–82.CrossRefPubMed
53.
Zurück zum Zitat Ougolkov AV, Bone ND, Fernández-Zapico ME, Kay NE, Billadeau DD. Inhibition of glycogen synthase kinase-3 activity leads to epigenetic silencing of nuclear factor kappaB target genes and induction of apoptosis in chronic lymphocytic leukemia B cells. Blood. 2007;110:735–42.CrossRefPubMedPubMedCentral Ougolkov AV, Bone ND, Fernández-Zapico ME, Kay NE, Billadeau DD. Inhibition of glycogen synthase kinase-3 activity leads to epigenetic silencing of nuclear factor kappaB target genes and induction of apoptosis in chronic lymphocytic leukemia B cells. Blood. 2007;110:735–42.CrossRefPubMedPubMedCentral
54.
Zurück zum Zitat Zhou Y, Uddin S, Zimmerman T, Kang JA, Ulaszek J, Wickrema A. Growth control of multiple myeloma cells through inhibition of glycogen synthase kinase-3. Leuk Lymphoma. 2008;49:1945–53.CrossRefPubMedPubMedCentral Zhou Y, Uddin S, Zimmerman T, Kang JA, Ulaszek J, Wickrema A. Growth control of multiple myeloma cells through inhibition of glycogen synthase kinase-3. Leuk Lymphoma. 2008;49:1945–53.CrossRefPubMedPubMedCentral
55.
Zurück zum Zitat Song EY, Palladinetti P, Klamer G, Ko KH, Lindeman R, O¨Brien TA, et al. Glycogen synthase kinase–3β inhibitors suppress leukemia cell growth. Exp Hematol. 2010;38:908–21.CrossRefPubMed Song EY, Palladinetti P, Klamer G, Ko KH, Lindeman R, O¨Brien TA, et al. Glycogen synthase kinase–3β inhibitors suppress leukemia cell growth. Exp Hematol. 2010;38:908–21.CrossRefPubMed
56.
Zurück zum Zitat Banerji V, Frumm SM, Ross KN, Li LS, Schinzel AC, Hahn CK, et al. The intersection of genetic and chemical genomic screens indentifies GSK-3α as a target in human acute meyloid leukemia. J Clin Invest. 2012;122:935–47.CrossRefPubMedPubMedCentral Banerji V, Frumm SM, Ross KN, Li LS, Schinzel AC, Hahn CK, et al. The intersection of genetic and chemical genomic screens indentifies GSK-3α as a target in human acute meyloid leukemia. J Clin Invest. 2012;122:935–47.CrossRefPubMedPubMedCentral
Metadaten
Titel
Selected polyphenols potentiate the apoptotic efficacy of glycolytic inhibitors in human acute myeloid leukemia cell lines. Regulation by protein kinase activities
verfasst von
Elena de Blas
María Cristina Estañ
María del Carmen Gómez de Frutos
Javier Ramos
María del Carmen Boyano-Adánez
Patricio Aller
Publikationsdatum
01.12.2016
Verlag
BioMed Central
Erschienen in
Cancer Cell International / Ausgabe 1/2016
Elektronische ISSN: 1475-2867
DOI
https://doi.org/10.1186/s12935-016-0345-y

Weitere Artikel der Ausgabe 1/2016

Cancer Cell International 1/2016 Zur Ausgabe

Update Onkologie

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.