ArticlePDF Available

Involvement of Hydrogen Peroxide in Safingol-Induced Endonuclease G-Mediated Apoptosis of Squamous Cell Carcinoma Cells

Authors:

Abstract and Figures

Safingol, a L-threo-dihydrosphingosine, induced the nuclear translocation of a mitochondrial apoptogenic mediator-endonuclease G (endo G)-and apoptosis of human oral squamous cell carcinoma (SCC) cells. Upstream mediators remain largely unknown. The levels of hydrogen peroxide (H2O2) in cultured oral SCC cells were measured. Treatment with safingol increased intracellular H2O2 levels but not extracellular H2O2 levels, indicating the production of H2O2. The cell killing effect of safingol and H2O2 was diminished in the presence of reactive oxygen species (ROS) scavenger N-acetyl-L-cysteine (NAC). Dual staining of cells with annexin V and propidium iodide (PI) revealed that apoptotic cell death occurred by treatment with H2O2 and safingol. The number of apoptotic cells was reduced in the presence of NAC. In untreated cells, endo G distributed in the cytoplasm and an association of endo G with mitochondria was observed. After treatment with H2O2 and safingol, endo G was distributed to the nucleus and cytoplasm, indicating the nuclear translocation of the mitochondrial factor. NAC prevented the increase of apoptotic cells and the translocation of endo G. Knock down of endo G diminished the cell killing effect of H2O2 and safingol. These results suggest that H2O2 is involved in the endo G-mediated apoptosis of oral SCC cells by safingol.
Content may be subject to copyright.
Int. J. Mol. Sci. 2014, 15, 2660-2671; doi:10.3390/ijms15022660
International Journal of
Molecular Sciences
ISSN 1422-0067
www.mdpi.com/journal/ijms
Article
Involvement of Hydrogen Peroxide in Safingol-Induced
Endonuclease G-Mediated Apoptosis of Squamous Cell
Carcinoma Cells
Masakazu Hamada *, Ken Wakabayashi, Atsushi Masui, Soichi Iwai, Tomoaki Imai and
Yoshiaki Yura
Department of Oral and Maxillofacial Surgery, Osaka University Graduate School of Dentistry,
1-8 Yamadaoka, Suita, Osaka 565-0871, Japan; E-Mails: dental_ken@yahoo.co.jp (K.W.);
a-masui@dent.osaka-u.ac.jp (A.M.); s-iwai@dent.osaka-u.ac.jp (S.I.); hsc12@hotmail.com (T.I.);
yura@dent.osaka-u.ac.jp (Y.Y.)
* Author to whom correspondence should be addressed; E-Mail: hmdmskz@dent.osaka-u.ac.jp;
Tel.: +81-6-6879-2941; Fax: +81-6-6879-2170.
Received: 8 November 2013; in revised form: 3 January 2014 / Accepted: 13 February 2014 /
Published: 17 February 2014
Abstract: Safingol, a L-threo-dihydrosphingosine, induced the nuclear translocation of a
mitochondrial apoptogenic mediator—endonuclease G (endo G)—and apoptosis of human
oral squamous cell carcinoma (SCC) cells. Upstream mediators remain largely unknown.
The levels of hydrogen peroxide (H
2
O
2
) in cultured oral SCC cells were measured.
Treatment with safingol increased intracellular H
2
O
2
levels but not extracellular H
2
O
2
levels,
indicating the production of H
2
O
2
. The cell killing effect of safingol and H
2
O
2
was
diminished in the presence of reactive oxygen species (ROS) scavenger N-acetyl-L-cysteine
(NAC). Dual staining of cells with annexin V and propidium iodide (PI) revealed that
apoptotic cell death occurred by treatment with H
2
O
2
and safingol. The number of apoptotic
cells was reduced in the presence of NAC. In untreated cells, endo G distributed in the
cytoplasm and an association of endo G with mitochondria was observed. After treatment
with H
2
O
2
and safingol, endo G was distributed to the nucleus and cytoplasm, indicating the
nuclear translocation of the mitochondrial factor. NAC prevented the increase of apoptotic
cells and the translocation of endo G. Knock down of endo G diminished the cell killing
effect of H
2
O
2
and safingol. These results suggest that H
2
O
2
is involved in the endo
G-mediated apoptosis of oral SCC cells by safingol.
OPEN ACCESS
Int. J. Mol. Sci. 2014, 15 2661
Keywords: safingol; hydrogen peroxide; apoptosis; endonuclease G
1. Introduction
Apoptosis, the best-described type of programmed cell death, is characterized by cell membrane
blebbing, a reduction in cellular volume, the activation of caspases, chromatin condensation and nuclear
fragmentation [1,2]. Internucleosomal DNA fragmentation is a hallmark of the apoptotic process and at
least two endonucleases, caspase-activated DNase (CAD) and endonuclease G (endo G), are thought to be
important for mammalian DNA fragmentation during apoptosis [3,4]. The best-characterized major
enzyme for DNA fragmentation is the CAD that forms an inactive heterodimer with inhibitor of CAD
(ICAD). Following apoptotic signaling, ICAD is proteolyzed by caspase-3 causing the dissociation of
the CAD/ICAD heterodimer and releasing CAD, which then moves from the cytosol to the nucleus.
Endo G is an endonuclease that is released from the mitochondrial intermembrane space and translocates
to the cell nucleus to induce DNA fragmentation in a caspase-independent manner [4–6].
Safingol, a L-threo-dihydrosphingosine, is a synthetic lipid and functions by targeting the
lipid-binding regulatory domain of protein kinase C (PKC) [7,8]. In previous studies, safingol was used
as a PKCα-selective inhibitor and antitumor activity was demonstrated [8–12]. The cytotoxic effect of
safingol was also attributed to the inhibition of sphingosine kinase 1, thus preventing the formation of
sphingosine-1-phosphate, which is involved in cell proliferation, invasion and angiogenesis [13–15].
Safingol is currently under a phase I clinical trial in combination with cisplatin for the treatment of
advanced solid tumors [16]. Our previous studies indicated that safingol induced apoptosis of oral
squamous cell carcinoma (SCC) cells, accompanied by the nuclear translocation of endo G from
mitochondria in a caspase 3-independet manner, using DNA fragmentation assay, flow cytometric
analysis and immunostaining [17], but upstream mediators remain largely unknown.
Oxidative stress has been implicated in a number of physiological and pathological processes,
including cancer, ischemic injury, neurodegenerative diseases, chronic inflammation, type II diabetes
and arteriosclerosis [18]. Reactive oxygen species (ROS) are recognized as chemical mediators in
deciding the fate of cells, depending on the extent of oxidative damage. In the present study, we
investigated the possible involvement of H
2
O
2
as a ROS in endo G-mediated apoptosis of oral SCC cells
treated with safingol.
2. Results
2.1. Production of ROS in SCC Cells by Treatment with Safingol
SAS cells were incubated with hydrogen peroxide (H
2
O
2
) or safingol, and extracellular and
intracellular levels of H
2
O
2
were measured using an assay kit for measuring H
2
O
2
concentration [19,20]
12 h later. After treatment with 100 µM H
2
O
2
, the intracellular H
2
O
2
concentration increased, but the
extracellular H
2
O
2
concentration did not (Figure 1A). When SAS cells were treated with safingol at
15 or 25 µM, the H
2
O
2
levels in the cells also increased. The difference between the treated cells and
untreated control was significant. The level of H
2
O
2
in the medium of the cells treated with H
2
O
2
or
Int. J. Mol. Sci. 2014, 15 2662
safingol was not altered. When SAS cells were treated with 15 µM safingol for 6, 12, and 24 h, the
intracellular H
2
O
2
concentration increased and reached a max level of 12 h.
Figure 1. Production of ROS in SCC cells treated with safingol. SAS cells were treated with
H
2
O
2
or safingol, and extracellular and intracellular ROS levels were determined 12 h later
(A); SAS cells were treated with 15 µM safingol, and intracellular ROS levels were
determined 0, 6, 12, 24 h later (B). The data represent the mean ± SD of three
determinations. ** p < 0.01 vs. control.
2.2. Induction of Cell Death by H
2
O
2
and Safingol
Cell death was examined using the trypan blue dye exclusion test. When SAS cells were treated with
100 µM H
2
O
2
for 12 h, the proportion of dead cells increased to 36%, though this increase was
diminished in the presence of a ROS scavenger, N-acetyl-L-cysteine (NAC) [21,22], with 22% of cells
nonviable (Figure 2A). When cells were treated with 15 µM safingol for 12 h, 45% were found to be
nonviable. This value was reduced to 21% by NAC. When another oral SCC cell line HSC-3 was used,
H
2
O
2
and safingol decreased the proportion of viable cells in a similar manner as observed in SAS cells.
The suppressive effect was blunted by NAC (Figure 2B). When 500U PEG catalase (PEG-cat) was used
to delete H
2
O
2
production by safingol, the cell killing effect of safingol was decreased (Figure 2C).
SAS cells were treated with H
2
O
2
or safingol and dual staining with annexin V and propidium iodide
(PI) was performed. Cells stained with annexin V alone were considered to be apoptotic cells.
The percentage of apoptotic cells was also increased by treatment with H
2
O
2
and safingol, up to 34% and
23%, respectively (Figure 3). These values decreased to 14% and 15% in the presence of NAC.
Int. J. Mol. Sci. 2014, 15 2663
Figure 2. Induction of cell death by H
2
O
2
and safingol. SAS (A) and HSC-3 (B) cells were
treated with 100 µM H
2
O
2
or 15 µM safingol alone. Alternatively, they were treated with
H
2
O
2
or safingol in the presence of the ROS scavenger NAC for 12 h. SAS cells were
treated with 100 µM H
2
O
2
or 15 µM safingol alone. Alternatively, they were pretreated
with 500U PEG catalase (PEG-cat) for 2 h and then they were treated with H
2
O
2
or
safingol for 12 h (C). Thereafter, the cells were stained with trypan blue. The percentages
of dead cells were calculated. The data represent the mean ± SD of three determinations.
* p < 0.05, ** p < 0.01 vs. treated group with NAC or PEG-cat.
Figure 3. Induction of apoptotic cells death by H
2
O
2
and safingol. SAS cells were treated
with 100 µM H
2
O
2
or 15 µM safingol alone (A). Alternatively, they were also treated with
H
2
O
2
or safingol in the presence of the ROS scavenger NAC for 12 h. Thereafter, the cells
were stained with annexin V and PI. The percentages of apoptotic cells stained with annexin
V alone were calculated (B). The data represent the mean ± SD of three determinations.
** p < 0.01 vs. treated group with NAC.
2.3. Effect of Endo G Small Interfering RNA (siRNA) on the Cell Death Caused by H
2
O
2
and Safingol
Previously, we reported that safingol induced the translocation of endo G from mitochondria to the
nucleus and induced apoptosis [17]. In the present study, the effect of siRNA on cell viability was
examined. SAS cells were transfected with endo G siRNA and subjected to immunoblotting. The
expression of endo G was downregulated by this treatment, whereas it was maintained after the
transfection of nonsense siRNA (Figure 4A).
Int. J. Mol. Sci. 2014, 15 2664
Treatment with 300 µM H
2
O
2
and 15 µM safingol increased the percentage of dead cells to 33% and
27%, respectively, in the cultures transfected with nonsense siRNA. In endo G siRNA-transfected cells,
these values decreased to 14% and 17%, respectively, indicating the involvement of endo G in the H
2
O
2
-
and safingol-induced cell death (Figure 4B).
Figure 4. Effect of endo G siRNA transfection on cell viability. (A) SAS cells were
transfected with endo G siRNA or nonsense siRNA and cultured for 24 h. They were
subjected to an immunoblot analysis. At least three determinations were performed.
A representative result is shown; (B) Endo G siRNA- or nonsense siRNA-transfected SAS
cells were treated with 300 µM H
2
O
2
or 15 µM safingol for 12 h and subjected to a trypan
blue dye exclusion test. The data represent the mean ± SD of three determinations.
** p < 0.01 vs. treated group with endo G siRNA transfection.
2.4. Translocation of Endo G by H
2
O
2
and Safingol
The effect of H
2
O
2
and safingol on the localization of mitochondria and the expression of endo G
were examined using immunofluoresent antibody staining. In untreated SAS cells, the mitochondria
were filamentous with a tubular appearance and often interconnected forming a network. Most
cytoplasmic staining of endo G was co-localized with mitochondria, and specific nuclear staining was
not observed (Figure 5). After treatment with 100 µM H
2
O
2
, the localization of mitochondria
was unchanged, but endo G showed diffuse distribution. Inconsistent with the results for
4',6-diamidino-2-phenylindole (DAPI) staining, nuclear staining of endo G was observed. When cells
were treated with H
2
O
2
in the presence of NAC, the endo G was confined to the cytoplasm and nuclear
localization was not observed. Safingol also induced nuclear staining of endo G, which was completely
blocked by NAC.
Int. J. Mol. Sci. 2014, 15 2665
Figure 5. Translocation of endo G by H
2
O
2
and safingol. SAS cells were treated with
100 µM H
2
O
2
or 15 µM safingol in the presence or absence of NAC for 12 h. They were
subjected to immunofluorescent staining using DAPI, antibody against endo G and
Mitotracker Red CMXRos DAPI. Untreated cells were also stained. The imaging of each
staining was merged. At least three measurements were performed. A representative result
is shown.
3. Discussion
We had shown that safingol could induce cell death with characteristics of apoptosis at a
concentration of 25 µM in a caspase three-independent manner [23]. At 10 µM, however, though a
proportion of cells detached, they reattached on the plate after prolonged incubation. The cell killing
effect of safingol was marginal at this concentration [17,23]. On the other hand, safingol was reported to
exert an inhibitory effect on sphingosine kinase 1 at concentrations below 10 µM [14,16]. Since our
previous study was undertaken at higher concentrations, safingol would affect sphingosine kinase 1 as
well as PKC, to induce apoptosis of oral SCC cells.
Activation of apoptosis is associated with the generation of ROS [24]. Indeed, some anticancer drugs
induce production of ROS during apoptosis [25–27]. Mizutani et al. [27] reported that the critical
apoptotic trigger of doxorubicin, a topoisomerase II inhibitor, was oxidative DNA damage from
doxorubicin-induced H
2
O
2
production and that the oxidative damage caused the indirect generation of
H
2
O
2
through poly (ADP-ribose) polymerase (PARP) and nicotinamide adenine dinucleotide phosphate
(NADPH) oxidase activation, leading to an increase in mitochondrial membrane permeability.
To determine the possible involvement of ROS in safingol-induced cell death in oral SCC cells, we
Int. J. Mol. Sci. 2014, 15 2666
examined the levels of intracellular and extracellular ROS and found that safingol as well as H
2
O
2
increased intracellular H
2
O
2
level. When cell death was estimated using the trypan blue dye exclusion
test, H
2
O
2
and safingol induced cell death and the killing effect was efficiently blocked by a ROS
scavenger, NAC, in SAS and HSC-3 cells. Apoptotic cells stained with annexin V alone were also
reduced by NAC in SAS cells. The difference observed in trypan blue and annexin V staining especially
in safingol-treated cells may represent apoptotic cells with necrotic degradation, because apoptotic cells
looks like necrotic cells at an advanced stage. Thus, it can be stated that safingol produces ROS
including H
2
O
2
, which is responsible for the induction of apoptotic cell death in oral SCC cells.
Mitochondria itself produces ROS, but H
2
O
2
that was added to the cell culture induced apoptosis in the
present study. Cytoplasmic H
2
O
2
must be the inducer of release of apoptogenic mitochondrial factors.
Ling et al. [28] indicated that safingol caused time-and concentration-dependent production of ROS in
MDA-MB-231 and HT-29 cells, suggesting ROS to be a mediator of safingol-induced cancer cell death.
They cultured the cells for 48 h at 10 µM and found necrosis and autophagy, but they did not examine
DNA fragmentation observed in apoptosis. Since safingol at 10 µM did not induce cell death in most
SAS cells, we have not examined the role of autophagy as observed in MDA-MB-231 and HT-29 cells.
Intrinsic apoptosis is critically dependent on mitochondrial outer membrane permeabilization,
which results in the release of mitochondrial intermembrane space proteins, such as cytochrome c, and
endo G [29,30]. In the present study, we found that the effect of H
2
O
2
and safingol was blocked by the
downregulation of endo G expression, indicating that endo G is required for the cell death by the
treatment. We also found that H
2
O
2
as well as safingol induced translocation of endo G to the nucleus.
The expression of endo G was not necessarily correlated with the intensity of mitochondrial staining, but
was present in the cytoplasm of untreated oral SCC cells, representing the synthesis of the mitochondrial
protein in the cytoplasm. After treatment with H
2
O
2
or safingol, the nuclear accumulation of endo G
occurred, but cytoplasmic staining was preserved. Apoptotic signal would stimulate the release of endo
G from mitochondria to the cytoplasm and then the endo G comcomitantly with the preexisting
cytoplasmic endo G move to the nucleus for DNA fragmentaion. It should be also stated that
NAC clearly blocked the alteration of endo G staining by H
2
O
2
and safingol. In a neuronal system,
Higgins et al. [31] found that oxidative stress triggered neuronal caspase-independent cell death and the
translocation of endo G. Treatment caused the redistribution from mitochondria of both endo G and
cytochrome c. Kim et al. [32] treated head and neck cancer cells with cisplatin and found mitochondrial
outer membrane permeabilization, the nuclear translocation of endo G and apoptosis. Together,
we firstly suggest that the expression and translocation of endo G are required for the inducation of cell
death of oral SCC cells by safingol and that H
2
O
2
is one of the upstream factors in this event.
4. Experimental Section
4.1. Cell Culture
The human oral SCC cell line SAS and HSC-3 were obtained from the Japanese Collection of
Research Bioresources (Tokyo, Japan). Cells were cultured in Dulbecco’s modified Eagle’s medium
(DMEM) supplemented with 5% fetal bovine serum, 2 mM L-glutamine, 100 µg/mL penicillin and
100 µg/mL streptomycin and grown in an incubator at 37 °C in a humidified atmosphere with 5% CO
2
.
Int. J. Mol. Sci. 2014, 15 2667
4.2. Reagents
Safingol was obtained from Calbiochem-Novabiochem (San Diego, CA, USA). H
2
O
2
and NAC were
obtained from Wako (Osaka, Japan) and PEG-cat was obtained from Sigma (St. Louis, MO, USA).
4.3. Measurement of H
2
O
2
The concentration of H
2
O
2
was determined using a colorimetric assay. Cells were plated in 48-well
plates at a density of 2 × 10
4
cells/well and treated with H
2
O
2
or safingol for 12 h. The supernatant of SAS
cells was harvested as an extracellular sample. Cells were dissociated with an ethylenediaminetetraacetic
acid (EDTA)-trypsin solution, subjected to three cycles of freezing and thawing and used as an
intracellular sample for the H
2
O
2
assay [19,20]. Ten microliters of sample was mixed with 100 µL of
Bioxytech H
2
O
2
-560 (OXIS International, Portland, OR, USA) and incubated for 30 min at room
temperature. Measurements were made using a Benchmark plus microplate spectrophotometer
(Bio-Ras Laboratories, Hercules, CA, USA) at a wavelength of 560 nm.
4.4. Trypan Blue Staining
Cell viability was determined by the trypan blue dye exclusion test. Cells were plated in 6-well plates
at a density of 1 × 10
6
cells/well, cultured for 24 h and treated with 100 µM H
2
O
2
or 15 µM safingol for
12 h. They were dissociated by the EDTA-trypsin solution, and cells were centrifuged suspended in
phosphate-buffered saline (PBS) without Ca
2+
and Mg
2+
. The pellets were then mixed with an equal
volume of PBS without Ca
2+
and Mg
2+
containing 0.4% trypan blue and observed with a microscope.
We counted the numbers of stained and unstained cells. Results were compared to those for the
untreated controls and a percentage was calculated.
4.5. Annexin V and PI Staining
To identify apoptotic cells, annexin V and PI staining was performed using Vybrant Apoptosis Assay
Kit (Life Technologies Corporation, Carlsbad, CA, USA) according to the manufacturer’s directions.
After treatment with H
2
O
2
or safingol, floating cells were harvested with medium and attached cells
were dissociated with EDTA-trypsin solution. These cells were collected by centrifugation at 1000 rpm
for 5 min. The cell pellets were suspended in 100 µL binding buffer (10 mM HEPES, 140 mM NaCl,
2.5 mM CaCl
2
, pH 7.4) and incubated with 5 µL FITC Annexin V and 1 µL of a propidium iodide
(100 µg/mL) solution for 15 min at room temperature. Staining for annexin V and propidium iodide was
observed under a fluorescence microscope (Microphoto FXA; Nikon, Tokyo, Japan). The percentages of
apoptotic cells stained with annexin V alone were calculated. At least 3 samples and 1000 cells were
counted for determination of the percentage of apoptotic cells.
4.6. Immunoblot Analysis
Cells were washed in PBS and lysed in a buffer containing 20 mM Tris-HCl (pH 7.4), 0.1% SDS,
1% TritonX-100, 1% sodium deoxycholate and protease inhibitor cocktail. After sonication on ice and
subsequent centrifugation at 15,000× g for 10 min at 4 °C, the supernatant was collected and the
Int. J. Mol. Sci. 2014, 15 2668
protein concentration was determined using a Protein Assay Kit (Bio-Rad, Hercules, CA, USA).
Sample protein (15 µg) was electrophoresed through a polyacrylamide gel and transferred to a
polyvinylidene fluoride membrane (Millipore, Bedford, MA, USA) by electroblotting. The membrane
was probed with antibodies and antibody-binding was detected using an enhanced chemiluminescence
kit (GE Healthcare, Amersham, Buckinghamshire, UK) according to the manufacturer’s instructions.
The antibodies used were a rabbit polyclonal antibody against endo G (Sigma, St. Louis, MO, USA),
and β-actin (Sigma, St. Louis, MO, USA). The secondary antibodies used were horseradish
peroxidase-conjugated anti-rabbit IgG (Cell Signaling Technology, Beverly, MA, USA) and
peroxidase-conjugated anti-mouse IgG (Sigma, St. Louis, MO, USA).
4.7. siRNA Transfection
Chemically synthetic siRNA against endo G and AllStars negative control siRNA (nonsense siRNA)
were purchased from Qiagen (Valencia, CA, USA). The target sequence of the siRNA for endo G was
5'-AAAUGCCUGGAACAACCUUGA-3'. Cells were plated in 6-well plates at a density of
1 × 10
5
cells/well, cultured for 24 h, and transfected with 40 nM endo G siRNA or nonsense siRNA
using Lipofectamine 2000 (Invitrogen, Carlsbad, CA, USA) according to the manufacturer’s
directions. The medium was replaced with DMEM after 3 h and cells were used for experiments at
24 h after transfection.
4.8. Confocal Laser Microscopic Analysis
Cells were treated with H
2
O
2
or safingol for 12 h. Thereafter, they were fixed with 2%
paraformaldehyde, permealized with 0.1% Triton X-100 in PBS and incubated with 25 nM
Mitotracker Red CMXRos (Molecular Probes, Eugene, OR, USA) at 37 °C for 45 min. They were
fixed in 4% paraformaldehyde phosphate buffer solution (WAKO, Osaka, Japan), permealized with
0.1% Triton X-100 in PBS and incubated with a rabbit polyclonal antibody against endo G (Sigma,
St. Louis, MO, USA) diluted 1:200 in PBS for 1 h at room temperature. After washing, the cells were
incubated with Alexa Fluor 488 goat antirabbit antibody (Life Technologies Corporation, Carlsbad,
CA, USA) diluted 1:500 in PBS for 1 h. After washing, coverslips were mounted onto microslides
using a ProLong Gold Antifade Reagent with DAPI (Life Technologies Corporation, Carlsbad, CA,
USA). The slides were analyzed under a confocal laser-scanning microscope Leica TCS SP8
(Leica Microsystems, Mannheim, Germany).
4.9. Statistical Analysis
The statistical analysis was performed using a Student’s t test with Microsoft Excel (Windows vista,
Microsoft, Redmond, WA, USA). The results were expressed as the mean ± SD. The differences were
considered significant at p < 0.05.
Int. J. Mol. Sci. 2014, 15 2669
5. Conclusions
Safingol mimics H
2
O
2
in the ability to induce the death of oral SCC cells. H
2
O
2
as a ROS is
suggested to act as upstream mediators to induce translocation of endo G which can directly contribute
to the cleavage of nuclear DNA.
Acknowledgments
This work was supported in part by a Grant-in Aid for Scientific Research from the Japan Society
for the Promotion of Science (No.22791972, No.25861929).
Conflicts of Interest
The authors declare no conflict of interests.
References
1. Kroemer, G.; Galluzzi, L.; Vandenabeele, P.; Abrams, J.; Alnemri, E.S.; Baehrecke, E.H.;
Blagosklonny, M.V.; El-Deiry, W.S.; Golstein, P.; Green, D.R.; et al. Classification of cell death:
recommendations of the Nomenclature Committee on Cell Death 2009. Cell. Death Differ. 2009,
16, 3–11.
2. Ghobrial, I.M.; Witzig, T.E.; Adjei, A.A. Targeting apoptosis pathways in cancer therapy.
CA: A Cancer J. Clin. 2005, 55, 178–194.
3. Sakahira, H.; Enari, M.; Nagata, S. Cleavage of CAD inhibitor in CAD activation and DNA
degradation during apoptosis. Nature 1998, 391, 96–99.
4. Li, L.Y.; Luo, X.; Wang, X. Endonuclease G is an apoptotic DNase when released from
mitochondria. Nature 2001, 412, 95–99.
5. Lorenzo, H.K.; Susin, S.A. Mitochondrial effectors in caspase-independent cell death. FEBS Lett.
2004, 557, 14–20.
6. Li, J.; Zhou, J.; Li, Y.; Qin, D.; Li, P. Mitochondrial fission controls DNA fragmentation by
regulating endonuclease G. Free Radic. Biol. Med. 2010, 49, 622–631.
7. Hannun, Y.A.; Loomis, C.R.; Merrill, A.H., Jr.; Bell, R.M. Sphingosine inhibition of protein kinase
C activity and of phorbol dibutyrate binding in vitro and in human platelets. J. Biol. Chem. 1986,
261, 12604–12609.
8. Kedderis, L.B.; Bozigian, H.P.; Kleeman, J.M.; Hall, R.L.; Palmer, T.E.; Harrison, S.D., Jr.;
Susick, R.L., Jr. Toxicity of the protein kinase C inhibitor safingol administered alone and in
combination with chemotherapeutic agents. Fundam. Appl. Toxicol. 1995, 25, 201–217.
9. Schwartz, G.K.; Haimovitz-Friedman, A.; Dhupar, S.K.; Ehleiter, D.; Maslak, P.; Lai, L.;
Loganzo, F., Jr.; Kelsen, D.P.; Fuks, Z.; Albino, A.P. Potentiation of apoptosis by treatment with
the protein kinase C-specific inhibitor safingol in mitomycin C-treated gastric cancer cells.
J. Natl. Cancer Inst. 1995, 87, 1394–1399.
10. Hoffmann, T.K.; Leenen, K.; Hafner, D.; Balz, V.; Gerharz, C.D.; Grund, A.; Ballo, H.;
Hauser, U.; Bier, H. Antitumor activity of protein kinase C inhibitors and cisplatin in human head
and neck squamous cell carcinoma lines. Anticancer Drugs 2002, 13, 93–100.
Int. J. Mol. Sci. 2014, 15 2670
11. Choe, Y.; Jung, H.; Khang, I.; Kim, K. Selective roles of protein kinase C isoforms on cell motility
of GT1 immortalized hypothalamic neurones. J. Neuroendocrinol. 2003, 15, 508–515.
12. Uemura, K.; Aki, T.; Yamaguchi, K.; Yoshida, K. Protein kinase C-epsilon protects PC12 cells
against methamphetamine-induced death: Possible involvement of suppression of glutamate
receptor. Life Sci. 2003, 72, 1595–1607.
13. Buehrer, B.M.; Bell, R.M. Sphingosine kinase: Properties and cellular functions. Adv. Lipid Res.
1993, 26, 59–67.
14. Olivera, A.; Kohama, T.; Tu, Z.; Milstien, S.; Spiegel, S. Purification and characterization of rat
kidney sphingosine kinase. J. Biol. Chem. 1998, 273, 12576–12583.
15. Pyne, N.J.; Pyne, S. Sphingosine 1-phosphate and cancer. Nat. Rev. Cancer 2010, 10, 489–503.
16. Dickson, M.A.; Carvajal, R.D.; Merrill, A.H., Jr.; Gonen, M.; Cane, L.M.; Schwartz, G.K. A phase
I clinical trial of safingol in combination with cisplatin in advanced solid tumors.
Clin. Cancer Res. 2011, 17, 2484–2492.
17. Hamada, M.; Sumi, T.; Iwai, S.; Nakazawa, M.; Yura, Y. Induction of endonuclease G-mediated
apopotosis in human oral squamous cell carcinoma cells by protein kinase C inhibitor safingol.
Apoptosis 2006, 11, 47–56.
18. Azad, M.B.; Chen, Y.; Gibson, S.B. Regulation of autophagy by reactive oxygen species (ROS):
Implications for cancer progression and treatment. Antioxid Redox. Signal. 2009, 11, 777–790.
19. Jiang, Z.Y.; Woollard, A.C.; Wolff, S.P. Hydrogen peroxide production during experimental
protein glycation. FEBS Lett. 1990, 268, 69–71.
20. Jiang, Z.Y.; Hunt, J.V.; Wolff, S.P. Ferrous ion oxidation in the presence of xylenol orange for
detection of lipid hydroperoxide in low density lipoprotein. Anal. Biochem. 1992, 202, 384–389.
21. Zafarullah, M.; Li, W.Q.; Sylvester, J.; Ahmad, M. Molecular mechanisms of N-acetylcysteine
actions. Cell. Mol. Life Sci. 2003, 60, 6–20.
22. Aruoma, O.I.; Halliwell, B.; Hoey, B.M.; Butler, J. The antioxidant action of N-acetylcysteine: Its
reaction with hydrogen peroxide, hydroxyl radical, superoxide, and hypochlorous acid.
Free Radic. Biol. Med. 1989, 6, 593–597.
23. Noda, T.; Iwai, S.; Hamada, M.; Fujita, Y.; Yura, Y. Induction of apoptosis of detached oral
squamous cell carcinoma cells by safingol. Possible role of Bim, focal adhesion kinase and
endonuclease G. Apoptosis 2009, 14, 287–297.
24. Cai, J.; Jones, D.P. Superoxide in apoptosis. Mitochondrial generation triggered by cytochrome c
loss. J. Biol. Chem. 1998, 273, 11401–11404.
25. Gewirtz, D.A. A critical evaluation of the mechanisms of action proposed for the antitumor effects
of the anthracycline antibiotics adriamycin and daunorubicin. Biochem. Pharmacol. 1999, 57,
727–741.
26. Varbiro, G.; Veres, B.; Gallyas, F., Jr.; Sumegi, B. Direct effect of taxol on free radical formation
and mitochondrial permeability transition. Free Radic. Biol. Med. 2001, 31, 548–558.
27. Mizutani, H.; Tada-Oikawa, S.; Hiraku, Y.; Kojima, M.; Kawanishi, S. Mechanism of apoptosis
induced by doxorubicin through the generation of hydrogen peroxide. Life Sci. 2005, 76,
1439–1453.
28. Ling, L.U.; Tan, K.B.; Lin, H.; Chiu, G.N. The role of reactive oxygen species and autophagy in
safingol-induced cell death. Cell Death Dis. 2011, 2, e129.
Int. J. Mol. Sci. 2014, 15 2671
29. Zamzami, N.; Kroemer, G. Methods to measure membrane potential and permeability transition in
the mitochondria during apoptosis. Methods Mol. Biol. 2004, 282, 103–115.
30. Jourdain, A.; Martinou, J.C. Mitochondrial outer-membrane permeabilization and remodelling in
apoptosis. Int. J. Biochem. Cell Biol. 2009, 41, 1884–1889.
31. Higgins, G.C.; Beart, P.M.; Nagley, P. Oxidative stress triggers neuronal caspase-independent
death: endonuclease G involvement in programmed cell death-type III. Cell. Mol. Life Sci. 2009,
66, 2773–2787.
32. Kim, J.S.; Lee, J.H.; Jeong, W.W.; Choi, D.H.; Cha, H.J.; Kim, D.H.; Kwon, J.K.; Park, S.E.;
Park, J.H.; Cho, H.R.; et al. Reactive oxygen species-dependent EndoG release mediates
cisplatin-induced caspase-independent apoptosis in human head and neck squamous carcinoma
cells. Int. J. Cancer 2008, 122, 672–680.
© 2014 by the authors; licensee MDPI, Basel, Switzerland. This article is an open access article
distributed under the terms and conditions of the Creative Commons Attribution license
(http://creativecommons.org/licenses/by/3.0/).
... Suppression of cell death by the ROS scavenger NAC Bioactive lipids, such as ceramides, have been reported to induce reactive oxygen species (ROS) production. [23][24][25] The effect of a ROS scavenger, N-acetyl-L-cysteine (NAC), on PF-543 induced cell death was investigated. HSC-3 cells were pretreated with NAC and treated with PF-543 for 72 h, then analyzed using flow cytometry. ...
... We also reported ROS production by safingol in head and neck SCC cells. 25 In this study, we found that the removal of ROS by N-acetyl-L-cysteine (NAC) prevented the production of necrotic cells by PF-543. Therefore, it is likely that ROS generated by PF-543 is a mediator of PF-543induced cell death and autophagy. ...
Article
Full-text available
Sphingosine kinase 1 (SphK1) overexpressed in head and neck squamous cell carcinoma (SCC) regulates tumor growth. The effects of PF-543, a specific SphK1 inhibitor, on human SCC cells were examined. The proportion of viable cells after PF-543 treatment decreased in a time- and dose-dependent manner, and cell death occurred in SphK1-expressing SCC cells. Flow cytometry analysis revealed that PF-543 induced both necrosis and apoptosis. PF-543 also induced granular accumulation of LC3 and conversion from LC3-I to LC3-II, which was blocked by autophagy inhibitors, wortmannin, 3-methyladenine (3-MA), and bafilomycin A1. Treatment of head and neck SCC cells with autophagy inhibitors and PF-543 increased the proportion of cells with necrosis and apoptosis, indicating that autophagy acts to promote cell survival. Reactive oxygen species (ROS) scavenger reduced the cytotoxicity of PF-543. These results demonstrated that PF-543 induces apoptosis, necrosis, and autophagy in human head and neck SCC cells, and that autophagy antagonizes either necrosis or apoptosis.
... We previously reported that the treatment of SAS cells with 15 μM safingol produced ROS and also that the ROS scavenger, N-acetyl-L-cysteine (NAC), could prevent apoptosis, suggesting ROS as an upstream factor in the endoG-mediated pathway [25]. SAS cells were treated with safingol and 3-MA in the presence of NAC for 24 h and analyzed using flow cytometry. ...
... ROS may mediate the induction of apoptosis and/or autophagy in several types of cancer cells [26,38,43]. A previous study showed that ROS production was involved in the safingol induced-apoptosis of oral SCC cells [25]. In the present study, the ROS scavenger NAC prevented the induction of cell death caused by the combination of safingol and 3-MA. ...
Article
Full-text available
Safingol, L- threo-dihydrosphingosine, induces cell death in human oral squamous cell carcinoma (SCC) cells through an endonuclease G (endoG) -mediated pathway. We herein determined whether safingol induced apoptosis and autophagy in oral SCC cells. Safingol induced apoptotic cell death in oral SCC cells in a dose-dependent manner. In safingol-treated cells, microtubule-associated protein 1 light chain 3 (LC3)-I was changed to LC3-II and the cytoplasmic expression of LC3, amount of acidic vesicular organelles (AVOs) stained by acridine orange and autophagic vacuoles were increased, indicating the occurrence of autophagy. An inhibitor of autophagy, 3-methyladenine (3-MA), enhanced the suppressive effects of safingol on cell viability, and this was accompanied by an increase in the number of apoptotic cells and extent of nuclear fragmentation. The nuclear translocation of endoG was minimal at a low concentration of safingol, but markedly increased when combined with 3-MA. The suppressive effects of safingol and 3-MA on cell viability were reduced in endoG siRNA- transfected cells. The scavenging of reactive oxygen species (ROS) prevented cell death induced by the combinational treatment, whereas a pretreatment with a pan-caspase inhibitor z-VAD-fmk did not. These results indicated that safingol induced apoptosis and autophagy in SCC cells and that the suppression of autophagy by 3-MA enhanced apoptosis. Autophagy supports cell survival, but not cell death in the SCC cell system in which apoptosis occurs in an endoG-mediated manner.
... The inhibitory effect of safingol (2S,3S)-2-aminooctadecane-1,3-diol on SPHK1, PKC, PI3K, and glucose uptake, combined with the presence of reactive oxygen species (ROS), altered the ceramide-dihydroceramide balance and induced apoptosis and autophagy in cancer cells (Companioni et al. 2021). Preclinical studies focusing on the effect of safingol in SCC (Hamada et al. 2014), multiple myeloma (Tsukamoto et al. 2015), breast, colon (Ling et al. 2011), prostate cancers (Morales et al. 2007c), and isolated hepatocytes are being conducted (Carfagna et al. 1996). Moreover, other chemotherapeutic agents were added to potentiate its effect such as the co-administration of mitomycin C in gastric cancer (Schwartz et al. 1995) or irinotecan in colon cancers (Ling et al. 2009). ...
Chapter
Sphingolipids, particularly their two main bioactive metabolites ceramide and sphingosine-1-phosphate (S1P), play a key role in cancer death and survival. Ceramides induce cancer cell death through apoptosis, autophagy, or necroptosis. S1P on the other hand inhibits apoptosis and stimulates proliferation, migration, metastasis, and drug resistance via receptor-dependent or receptor-independent pathways. Modulating the cellular sphingolipidome, through targeting specific enzymes or metabolites, is emerging as a promising pharmacological intervention that could limit cancer progression and improve disease outcomes. In this chapter, we highlight new pharmacological tools that have the potential to modulate key sphingolipid enzymes and metabolites to be used in cancer treatment.
... Safingol also affects the balance of ceramide/dihydroceramide levels. The inhibitory effects on signaling, particularly on PKCϵ and PI3k, concomitant with the presence of ROS (67) synergize to induce apoptosis (decreased Bcl-2 levels and increased caspase cleavage) (59,60,(62)(63)(64)(65)68) and/or autophagy (63, 67) ( Figure 3). According to preclinical studies, the combination of safingol with conventional chemotherapy agents, such as doxorubicin (67), irinotecan (66), and mitomycin C (65), potentiates their effects, inducing apoptotic cell death and ROS production in different cell lines. ...
Article
Full-text available
Sphingolipids are an extensive class of lipids with different functions in the cell, ranging from proliferation to cell death. Sphingolipids are modified in multiple cancers and are responsible for tumor proliferation, progression, and metastasis. Several inhibitors or activators of sphingolipid signaling, such as fenretinide, safingol, ABC294640, ceramide nanoliposomes (CNLs), SKI-II, α-galactosylceramide, fingolimod, and sonepcizumab, have been described. The objective of this review was to analyze the results from preclinical and clinical trials of these drugs for the treatment of cancer. Sphingolipid-targeting drugs have been tested alone or in combination with chemotherapy, exhibiting antitumor activity alone and in synergism with chemotherapy in vitro and in vivo. As a consequence of treatments, the most frequent mechanism of cell death is apoptosis, followed by autophagy. Aslthough all these drugs have produced good results in preclinical studies of multiple cancers, the outcomes of clinical trials have not been similar. The most effective drugs are fenretinide and α-galactosylceramide (α-GalCer). In contrast, minor adverse effects restricted to a few subjects and hepatic toxicity have been observed in clinical trials of ABC294640 and safingol, respectively. In the case of CNLs, SKI-II, fingolimod and sonepcizumab there are some limitations and absence of enough clinical studies to demonstrate a benefit. The effectiveness or lack of a major therapeutic effect of sphingolipid modulation by some drugs as a cancer therapy and other aspects related to their mechanism of action are discussed in this review.
... In addition, Bcl-xL expression was decreased and Bax expression was increased, resulting in ROS-mediated necrotic cell death [140]. At the doses of 25-50 µM, safingol induced ROS-mediated apoptosis of human oral SCC cells [141]. During this process, Bcl-xL expression was decreased and mitochondrial endonuclease G was released into the cytoplasm, resulting in DNA fragmentation in the nucleus [142,143]. ...
Article
Full-text available
In the treatment of advanced head and neck squamous cell carcinoma (HNSCC), including oral SCC, radiotherapy is a commonly performed therapeutic modality. The combined use of radiotherapy with chemotherapy improves therapeutic effects, but it also increases adverse events. Ceramide, a central molecule in sphingolipid metabolism and signaling pathways, mediates antiproliferative responses, and its level increases in response to radiotherapy and chemotherapy. However, when ceramide is metabolized, prosurvival factors, such as sphingosine-1-phosphate (S1P), ceramide-1-phosphate (C1P), and glucosylceramide, are produced, reducing the antitumor effects of ceramide. The activities of ceramide- and sphingosine-metabolizing enzymes are also associated with radio- and chemo-resistance. Ceramide analogs and low molecular-weight compounds targeting these enzymes exert anticancer effects. Synthetic ceramides and a therapeutic approach using ultrasound have also been developed. Inhibitors of ceramide- and sphingosine-metabolizing enzymes and synthetic ceramides can function as sensitizers of radiotherapy and chemotherapy for HNSCC.
... Among the different proteins that participate in the various stages of apoptosis processes, EndoG (Endonuclease G) is released from the mitochondria in a pro-apototic Bcl-2 family-dependent and caspase-independent manner after which is translocated to the nucleus where it cleaves DNA into large fragments, likely due to cooperation with DNase I van Loo et al., 2001;Widlak et al., 2001). In addition, several studies have shown the role of EndoG in tumor growth inhibition (Hamada et al., 2014;Winnard et al., 2008;Yoshida et al., 2006). ...
Article
Aim Immunotoxins are proteins that consist of an antibody fragment linked to a toxin, used as agents for targeted therapy of cancers. Although the most potent immunotoxins are made from bacterial and plant toxins, obstacles which contribute to poor responses are immunogenicity in patients and rapid development of neutralizing antibodies. In the present study we proposed a new therapeutic immunotoxin for targeted cancer therapy of ROR1 expressing cancers: an anti ROR1 single chain fragment variable antibody (scFv)-endonuclease G (anti ROR1 scFv-EndoG). Methods The three-dimensional structure of anti ROR1 scFv-EndoG protein was modeled and structure validation tools were employed to confirm the accuracy and reliability of the developed model. In addition, stability and integrity of the model were assessed by molecular dynamic (MD) simulation. Results All results suggested the protein model to be acceptable and of good quality. Conclusions Anti-ROR1 scFv-EndoG would be expected to bind to the ROR1 extracellular domain by its scFv portion and selectively deliver non-immunogenic human endonuclease G enzyme as an end-stage apoptosis molecule into ROR1-expressing cancer cells and lead rapidly to apoptosis. We believe that anti ROR1 and other anti-tumor antigen scFv-EndoG forms may be helpful for cancer therapy.
... 116 Safingol induces apoptosis in OSCC by generating reactive oxygen species. 117 In addition to SphK1 inhibitors, ABC294640, a SphK2 selective inhibitor, suppresses cell proliferation of a number of cancer cell lines and reduces the in vivo tumor growth of mammary adenocarcinoma xenografts. 66 ABC294640 also exhibits synergistic effects with sorafenib, a multikinase inhibitor. ...
Article
Full-text available
Head and neck squamous cell carcinoma (HNSCC) is the sixth most frequent cancer type, with an annual incidence of approximately half a million people worldwide. It has a high recurrence rate and an extremely low survival rate. This is due to limited availability of effective therapies to reduce the rate of recurrence, resulting in high morbidity and mortality of patients with advanced stages of the disease. HNSCC often develops resistance to chemotherapy and targeted drug therapy. Thus, to overcome the problem of drug resistance, there is a need to explore novel drug targets. Sphingosine-1-phosphate (S1P) is a bioactive sphingolipid involved in inflammation, tumor progression, and angiogenesis. S1P is synthesized intracellularly by two sphingosine kinases (SphKs). It can be exported to the extracellular space, where it can activate a family of G-protein-coupled receptors. Alternatively, S1P can act as an intracellular second messenger. SphK1 regulates tumor progression, invasion, metastasis, and chemoresistance in HNSCC. SphK1 expression is highly elevated in advanced stage HNSCC tumors and correlates with poor survival. In this article, we review current knowledge regarding the role of S1P receptors and enzymes of S1P metabolism in HNSCC carcinogenesis. Furthermore, we summarize the current perspectives on therapeutic approaches for targeting S1P pathway for treating HNSCC.
Article
Toxicological analysis relies on the measurement of cell death. DNA fragmentation is the preferential method in cell death assessment because it measures irreversible stage of cell death, and it is universal to all mechanisms of cell death. DNA fragmentation is provided by a group of nine cytotoxic DNases/endonucleases, cumulative activity of which is sufficient to eventually destroy all cellular DNAs to potentially reusable mononucleotides. Of all the methods for measuring DNA fragmentation, the terminal deoxynucleotidyl transferase dUTP nick end labeling (TUNEL) assay seems to be the best in satisfying the toxicology assessment because it is simple, quantitative, based on individual cell measurements, and is not destructive to the cells. The latter allows it to combine with quantitative immunocytochemistry and other histology and genetic techniques to perform mechanistic studies.
Article
Full-text available
Sphingosine kinase catalyzes the formation of the bioactive sphingolipid metabolite sphingosine 1-phosphate, which plays important roles in numerous physiological processes, including growth, survival, and motility. We have purified rat kidney sphingosine kinase 6 x 10(5)-fold to apparent homogeneity. The purification procedure involved ammonium sulfate precipitation followed by chromatography on an anion exchange column. Partially purified sphingosine kinase was found to be stabilized by the presence of high salt, and thus, a scheme was developed to purify sphingosine kinase using sequential dye-ligand chromatography steps (since the enzyme bound to these matrices even in the presence of salt) followed by EAH-Sepharose chromatography. This 385-fold purified sphingosine kinase bound tightly to calmodulin-Sepharose and could be eluted in high yield with EGTA in the presence of 1 M NaCl. After concentration, the calmodulin eluate was further purified by successive high pressure liquid chromatography separations on hydroxylapatite, Mono Q, and Superdex 75 gel filtration columns. Purified sphingosine kinase has an apparent molecular mass of approximately 49 kDa under denaturing conditions on SDS-polyacrylamide gel, which is similar to the molecular mass determined by gel filtration, suggesting that the active form is a monomer. Sphingosine kinase shows substrate specificity for D-erythro-sphingosine and does not catalyze the phosphorylation of phosphatidylinositol, diacylglycerol, ceramide, DL-threo-dihydrosphingosine, or N,N-dimethylsphingosine. However, the latter two sphingolipids were potent competitive inhibitors. With sphingosine as substrate, the enzyme had a broad pH optimum of 6.6-7.5 and showed Michaelis-Menten kinetics, with Km values of 5 and 93 microM for sphingosine and ATP, respectively. This study provides the basis for molecular characterization of a key enzyme in sphingolipid signaling.
Article
Full-text available
Safingol is a sphingolipid with promising anticancer potential, which is currently in phase I clinical trial. Yet, the underlying mechanisms of its action remain largely unknown. We reported here that safingol-induced primarily accidental necrotic cell death in MDA-MB-231 and HT-29 cells, as shown by the increase in the percentage of cells stained positive for 7-aminoactinomycin D, collapse of mitochondria membrane potential and depletion of intracellular ATP. Importantly, safingol treatment produced time- and concentration-dependent reactive oxygen species (ROS) generation. Autophagy was triggered following safingol treatment, as reflected by the formation of autophagosomes, acidic vacuoles, increased light chain 3-II and Atg biomarkers expression. Interestingly, scavenging ROS with N-acetyl-L-cysteine could prevent the autophagic features and reverse safingol-induced necrosis. Our data also suggested that autophagy was a cell repair mechanism, as suppression of autophagy by 3-methyladenine or bafilomycin A1 significantly augmented cell death on 2-5 μM safingol treatment. In addition, Bcl-xL and Bax might be involved in the regulation of safingol-induced autophagy. Finally, glucose uptake was shown to be inhibited by safingol treatment, which was associated with an increase in p-AMPK expression. Taken together, our data suggested that ROS was the mediator of safingol-induced cancer cell death, and autophagy is likely to be a mechanism triggered to repair damages from ROS generation on safingol treatment.
Article
Toxicity of the Protein Kinase C Inhibitor Safingol Administered Alone and in Combination with Chemotherapeutic Agents. Kedderis, L. B., Bozigian, H. P., Kleeman, J. M., Hall, R. L., Palmer, T. E., Harrison, S. D., Jr., and Susick, R. L., Jr. (1995). Fundam. Appl. Toxicol. 25, 201-217.
Article
To elucidate the potential role of mitochondria in Taxol-induced cytotoxicity, we studied its direct mitochondrial effects. In Percoll-gradient purified liver mitochondria, Taxol induced large amplitude swelling in a concentration-dependent manner in the μM range. Opening of the permeability pore was also confirmed by the access of mitochondrial matrix enzymes for membrane impermeable substrates in Taxol-treated mitochondria. Taxol induced the dissipation of mitochondrial membrane potential (ΔΨ) determined by Rhodamine123 release and induced the release of cytochrome c from the intermembrane space. All these effects were inhibited by 2.5 μM cyclosporine A. Taxol significantly increased the formation of reactive oxygen species (ROS) in both the aqueous and the lipid phase as determined by dihydrorhodamine123 and resorufin derivative. Cytochrome oxidase inhibitor CN−, azide, and NO abrogated the Taxol-induced mitochondrial ROS formation while inhibitors of the other respiratory complexes and cyclosporine A had no effect. We confirmed that the Taxol-induced collapse of ΔΨ and the induction of ROS production occurs in BRL-3A cells. In conclusion, Taxol-induced adenine nucleotide translocase-cyclophilin complex mediated permeability transition, and cytochrome oxidase mediated ROS production. Because both cytochrome c release and mitochondrial ROS production can induce suicide pathways, the direct mitochondrial effects of Taxol may contribute to its cytotoxicity.
Article
Sphingosine 1-phosphate (S1P) is an important mediator of cancer cell growth and proliferation. Production of S1P is catalyzed by sphingosine kinase 1 (SphK). Safingol, (l-threo-dihydrosphingosine) is a putative inhibitor of SphK. We conducted a phase I trial of safingol (S) alone and in combination with cisplatin (C). A 3 + 3 dose escalation was used. For safety, S was given alone 1 week before the combination. S + C were then administered every 3 weeks. S was given over 60 to 120 minutes, depending on dose. Sixty minutes later, C was given over 60 minutes. The C dose of 75 mg/m(2) was reduced in cohort 4 to 60 mg/m(2) due to excessive fatigue. Forty-three patients were treated, 41 were evaluable for toxicity, and 37 for response. The maximum tolerated dose (MTD) was S 840 mg/m(2) over 120 minutes C 60 mg/m(2), every 3 weeks. Dose-limiting toxicity (DLT) attributed to cisplatin included fatigue and hyponatremia. DLT from S was hepatic enzyme elevation. S pharmacokinetic parameters were linear throughout the dose range with no significant interaction with C. Patients treated at or near the MTD achieved S levels of more than 20 μmol/L and maintained levels greater than and equal to 5 μmol/L for 4 hours. The best response was stable disease in 6 patients for on average 3.3 months (range 1.8-7.2 m). One patient with adrenal cortical cancer had significant regression of liver and lung metastases and another had prolonged stable disease. S was associated with a dose-dependent reduction in S1P in plasma. Safingol, the first putative SphK inhibitor to enter clinical trials, can be safely administered in combination with cisplatin. Reversible dose-dependent hepatic toxicity was seen, as expected from preclinical data. Target inhibition was achieved with downregulation of S1P. The recommended phase II dose is S 840 mg/m(2) and C 60 mg/m(2), every 3 weeks.
Article
Mitochondria constantly undergo fusion and fission that are necessary for the maintenance of organelle fidelity. However, growing evidence has shown that abnormal mitochondrial fusion and fission participate in the regulation of apoptosis. Mitochondrial fusion is able to inhibit apoptosis, whereas mitochondrial fission is involved in the initiation of apoptosis. It remains elusive as to whether mitochondrial fission can regulate DNA fragmentation during apoptosis. Mitochondrial fission is triggered by dynamin-related protein-1 (Drp1), whereas mitofusin 1 (Mfn 1) is able to induce mitochondrial fusion. Here, we report that Drp1 is required for the release of endonuclease G from mitochondria. Knockdown of Drp1 can attenuate DNA fragmentation. Our data further show that Mfn 1 prevents endonuclease G release from mitochondria and the consequent DNA fragmentation. Intriguingly, Mfn 1 could inhibit the activation of caspase-3 and caspase-9, which are necessary for endonuclease G translocation to the nucleus. Our results provide novel evidence that DNA fragmentation is regulated by the mitochondrial fission machinery.
Article
There is substantial evidence that sphingosine 1-phosphate (S1P) is involved in cancer. S1P regulates processes such as inflammation, which can drive tumorigenesis; neovascularization, which provides cancer cells with nutrients and oxygen; and cell growth and survival. This occurs at multiple levels and involves S1P receptors, sphingosine kinases, S1P phosphatases and S1P lyase. This Review summarizes current research findings and examines the potential for new therapeutics designed to alter S1P signalling and function in cancer.
Article
Safingol [(2S,3S)-2-amino-1,3-octadecanediol] potentiates the toxicity of doxorubicin (DOX) and cisplatin (CIS) against tumor cells in vitro and in vivo . The present studies were conducted in rats and dogs to evaluate safingol toxicity when administered iv as a single agent and to evaluate safingol's ability to potentiate the toxicity of established chemotherapeutic agents to normal tissues in vivo . In an escalating dose study, dogs were administered safingol iv at 5, 10, 20, 30, 40, and 75 mg&sol;kg on Days 1 through 6. Necropsies were performed on Day 7. Red urine was observed at 10 mg&sol;kg and higher. Icterus was observed following 40 mg&sol;kg with additional signs of hypoactivity and anorexia occurring after 75 mg&sol;kg. Clinical and microscopic pathology revealed marked hepatotoxicity, venous degeneration and necrosis at injection sites, and evidence of intra-vascular hemolysis. Doses of 5, 20, or 40 mg safingol&sol;kg were utilized in single iv dose rat and dog studies. No evidence of adverse systemic toxicity was seen up to 20 mg&sol;kg in either species [for rats: C max &equals; 12,600 (males) or 17,133 (females) ng&sol;ml, AUC &equals; 3853 (males) or 4365 (females) ng × hr&sol;ml; for dogs: C max &equals; 2533 ng&sol;ml, AUC &equals; 2851 ng × hr&sol;ml (no sex differences)]. Local effects of venous irritation or intravascular hemolysis were observed at all doses in rats and at 20 and 40 mg&sol;kg in dogs. A dose of 40 mg&sol;kg [for rats: C max &equals; 31,233 (males) or 91,300 (females) ng&sol;ml, AUC &equals; 11,519 (males) or 18,620 (females) ng × hr&sol;ml; for dogs: C max &equals; 9033 ng&sol;ml, AUC &equals; 11,094 ng × hr&sol;ml (combined sex)] was associated with clinical pathologic and renal histomorphologic changes considered consequent to intravascular hemolysis in both species, lethality and testicular toxicity in rats, and clinical biochemical changes indicative of hepatobiliary injury in dogs. Studies indicated that hemolysis occurred during infusion, was not caused by circulating levels of safingol, and was a function of dose concentration and vein of delivery. Safingol at 10 or 20 mg&sol;kg was administered iv to rats 30–60 min prior to myelosuppressive iv doses of DOX, CIS, or cyclophosphamide (CYP). Hematology, plus renal function and morphology for CIS-treated animals, was assessed 4 and 14 days later. Safingol did not potentiate DOX-, CIS-, or CYP-mediated leukopenia&sol;thrombocytopenia. A minimal enhancement of CIS-mediated decrease in GFR and increase in creatinine was observed at 20 mg safingol&sol;kg. Dogs were administered 20 mg safingol&sol;kg iv followed 60 min later by 0.5 or 1.25 mg DOX&sol;kg or 0.75 or 2.0 mg CIS&sol;kg. A complete toxicologic assessment 4 and 29 days postdose failed to show potentiation of DOX toxicity by safingol or vice versa. A renal lesion was inferred in dogs administered 20 mg&sol;kg safingol and 2 mg&sol;kg CIS based on minimal to slight renal tubular regeneration observed 4 weeks post-treatment. There were no effects of safingol on the pharmacokinetic profiles of DOX or CIS or vice versa.
Article
To characterize neuronal death, primary cortical neurons (C57/Black 6 J mice) were exposed to hydrogen peroxide (H2O2) and staurosporine. Both caused cell shrinkage, nuclear condensation, DNA fragmentation and loss of plasma membrane integrity. Neither treatment induced caspase-7 activity, but caspase-3 was activated by staurosporine but not H2O2. Each treatment caused redistribution from mitochondria of both endonuclease G (Endo G) and cytochrome c. Neurons knocked down for Endo G expression using siRNA showed reduction in both nuclear condensation and DNA fragmentation after treatment with H2O2, but not staurosporine. Endo G suppression protected cells against H2O2-induced cell death, while staurosporine-induced death was merely delayed. We conclude that staurosporine induces apoptosis in these neurons, but severe oxidative stress leads to Endo G-dependent death, in the absence of caspase activation (programmed cell death-type III). Therefore, oxidative stress triggers in neurons a form of necrosis that is a systematic cellular response subject to molecular regulation.
Article
Many human pathologies are associated with defects in mitochondria such as diabetes, neurodegenerative diseases or cancer. This tiny organelle is involved in a plethora of processes in mammalian cells, including energy production, lipid metabolism and cell death. In the so-called intrinsic apoptotic pathway, the outer mitochondrial membrane (MOM) is premeabilized by the pro-apoptotic Bcl-2 members Bax and Bak, allowing the release of apoptogenic factors such as cytochrome c from the inter-membrane space into the cytosol. At the same time, mitochondria fragment in response to Drp-1 activation suggesting that mitochondrial fission could play a role in mitochondrial outer-membrane permeabilization (MOMP). In this review, we will discuss the link that could exist between mitochondrial fission and fusion machinery, Bcl-2 family members and MOMP.