ArticlePDF Available

Human T-cell lymphotropic virus type I-transformed T-cells have a partial defect in ceramide synthesis in response to N -(4-hydroxyphenyl)retinamide

Authors:

Figures

Content may be subject to copyright.
Biochem. J. (2005) 392, 231–239 (Printed in Great Britain) doi:10.1042/BJ20050578 231
Human T-cell lymphotropic virus type I-transformed T-cells have a partial
defect in ceramide synthesis in response to
N
-(4-hydroxyphenyl)retinamide
Nadine DARWICHE*1, Ghada ABOU-LTEIF*, Tarek NAJDI*, Lina KOZHAYA, Ahmad ABOU TAYYOUN*, Ali BAZARBACHI
and Ghassan S. DBAIBO†§1
*Department of Biology, American University of Beirut, Beirut, Lebanon, Department of Biochemistry, American University of Beirut, Beirut, Lebanon, Department of Internal Medicine,
American University of Beirut, Beirut, Lebanon, and §Department of Pediatrics, American University of Beirut, Beirut, Lebanon
Treatment with the synthetic retinoid HPR [N-(4-hydroxyphenyl)-
retinamide] causes growth arrest and apoptosis in HTLV-I (human
T-cell lymphotropic virus type-I)-positive and HTLV-I-negative
malignant T-cells [8]. It was observed that HPR-mediated growth
inhibition was associated with ceramide accumulation only in
HTLV-I-negative cells. The aim of the present study was to in-
vestigate the mechanism by which HPR differentially regulates
ceramide metabolism in HTLV-I-negative and HTLV-I-positive
malignant T-cells. Clinically achievable concentrations of HPR
caused early dose-dependent increases in ceramide levels only in
HTLV-I-negative cells and preceded HPR-induced growth sup-
pression. HPR induced de novo synthesis of ceramide in HTLV-I-
negative, but not in HTLV-I-positive, cells. Blocking ceramide
glucosylation in HTLV-I-positive cells, which leads to accumul-
ation of endogenous ceramide, rendered these cells more sensitive
to HPR. Exogenous cell-permeant ceramides that function par-
tially by generating endogenous ceramide induced growth sup-
pression in all tested malignant lymphocytes, were consist-
ently found to be less effective in HTLV-I-positive cells confirm-
ing their defect in de novo ceramide synthesis. Owing to its
multipotent activities, the HTLV-I-encoded Tax protein was
suspected to inhibit ceramide synthesis. Tax-transfected Molt-4
and HELA cells were less sensitive to HPR and C6-ceramide me-
diated growth inhibition respectively and produced lower levels
of endogenous ceramide. Together, these results indicate that
HTLV-I-positive cells are defective in de novo synthesis of cer-
amide and that therapeutic modalities that bypass this defect are
more likely to be successful.
Key words: ceramide, human T-cell lymphotropic virus type-I
(HTLV-1), N-(4-hydroxyphenyl)retinamide (HPR), Tax protein.
INTRODUCTION
The retinoid HPR [N-(4-hydroxyphenyl)retinamide] inhibits
growth and induces apoptosis in many human cell lines, including
many that are all-trans-retinoic acid-resistant [1]. HPR’s cyto-
toxicity is mediated through retinoic acid receptor-dependent and
-independent mechanisms [2], may involve p53-independent path-
ways [3,4], is associated with the generation of ROS (reactive oxy-
gen species) [5], is coupled with JNK (c-Jun N-terminal kinase)
activation [6] and may require elevated levels of ceramide [7].
We previously reported that at clinically achievable concen-
trations, HPR is a potent and selective inducer of G1cell cycle
arrest and apoptosis in HTLV-I-positive and HTLV-I-negative
malignant T-cells with no effect on normal T-lymphocytes [8].
In HTLV-I-negative cells only, HPR-induced apoptosis was as-
sociated with ceramide accumulation, a sharp decrease in mito-
chondrial membrane potential, and activation of caspases 8, 9
and 3, and could be partially reverted by the caspase inhibitor
z-VAD. In contrast, HTLV-I-positive cells had a slower apoptotic
response and required higher HPR concentrations suggesting that
Tax protein, or other HTLV-I products, protected infected cells
from ceramide accumulation and caspase-mediated apoptosis [8].
Ceramide, a sphingolipid secondary messenger molecule, has
been proposed as a co-ordinator of eukaryotic stress responses [9].
Many inducers of stress responses result in ceramide accumul-
ation, usually as a result of sphingomyelin breakdown or by
de novo synthesis, and sometimes as a result of inhibition of cer-
amide clearance through sphingomyelin synthase or ceramidases
[10,11].
Ceramide exerts growth suppressive effects including those on
differentiation [12], cell cycle arrest by dephosphorylation of Rb
protein [13,14], senescence by inhibiting telomerase [15] and
most importantly promotes apoptosis in a variety of cell types
[16,17]. Ceramide may induce apoptosis through ROS generation
[18], activation of JNK/stress-activated protein kinase [19] and
through both caspase-dependent and -independent mechanisms
[20]. Inducers of ceramide accumulation include TNF-α(tumour
necrosis factor-α), Fas ligand, interleukin-1, γ-interferon, CD28
ligation, complement, serum deprivation, γ-irradiation, heat
shock, ultraviolet radiation, and most chemotherapeutic agents
examined [21]. Defects in ceramide production render the cells
more resistant to killing by these inducers [22–24], while elevation
of endogenous ceramide levels lowers the threshold for apoptosis
induction by these agents [25,26]. Cell permeant ceramide ana-
logues can also produce the growth suppressive effects of these
inducers [27].
Several studies described elevations in ceramide levels fol-
lowing HPR treatment and its role in apoptosis [28]. HL-60
leukaemic cells were first shown to exhibit a transient rise in cer-
amide levels following HPR treatment. The addition of FB1
(fumonisin B1), an inhibitor of de novo ceramide synthesis, in-
hibited HPR-induced apoptosis, while the addition of caspase
inhibitors had no effect on ceramide levels suggesting that
ceramide synthesis occurs upstream of caspase activation [29].
Abbreviations used: Dh-C6,dihydro-C
6-ceramide; DMEM, Dulbecco’s modified Eagle’s medium; DGK, diacylglycerol kinase; FBS, fetal bovine serum;
FB1, fumonisin B1; GCS, glucosylceramide synthase; HPR,
N
-(4-hydroxyphenyl)retinamide; HTLV-I, human T-cell lymphotropic virus type-I; JNK, c-Jun
N-terminal kinase; NF-κB, nuclear factor-κB; PDMP, D,L-threo-1-phenyl-2-decanoylamino-3-morpholino-1-propanol; ROS, reactive oxygen species; TNF-α,
tumour necrosis factor-α.
1Correspondence should be addressed to either author (email darwichn@aub.edu.lb and gdbaibo@aub.edu.lb).
c
2005 Biochemical Society
232 N. Darwiche and others
Neuroblastoma cells showed a sustained elevation in ceramide
levels after HPR treatment, at concentrations that induced both
necrosis and apoptosis, and again FB1inhibited apoptosis [4].
This HPR-induced ceramide accumulation in neuroblastoma cell
lines was due to the co-ordinated activation of serine palmitoyl-
transferase and ceramide synthase in the de novo synthesis path-
way [30]. As a consequence, combinations of HPR and some
modulators of ceramide metabolism may provide a novel chemo-
therapeutic approach [31]. In previous studies, cytotoxic concen-
trations of HPR have been shown to increase ceramide levels in
acute lymphoblastic leukaemia cell lines, and inhibitors of the
ceramide de novo pathway abrogated this ceramide accumulation
[32]. Furthermore, PC-3 prostate cancer cells, which are relatively
resistant to HPR, were rendered much more sensitive by blocking
the metabolism of ceramide to glucosylceramide using tamoxifen
[33]. All of these studies firmly establish ceramide as an important
player in apoptosis induction by HPR.
In this report, we demonstrate that HPR induces distinct cer-
amide responses in HTLV-I-negative and HTLV-I-positive malig-
nant T-cells. Clinically achievable concentrations of HPR caused
early dose-dependent increases in ceramide levels only in
HTLV-I-negative cells, which preceded HPR-induced growth
suppression. Furthermore, HPR specifically induced de novo syn-
thesis of ceramide in these latter cells. Exogenous cell-permeant
ceramides induced growth suppression in all malignant lympho-
cytes tested. Blocking ceramide glucosylation in HTLV-I-positive
cells enhanced HPR cytotoxicity in these cells. Tax protein-
transfected cells were less sensitive to HPR-mediated growth
inhibition and generated lower levels of ceramide.
EXPERIMENTAL
Cell lines and culture conditions
The HTLV-I-transformed CD4+T-cell lines HuT-102, MT-2 and
C8166, and the HTLV-I-negative CD4+T-cell lines CEM, Jurkat
and Molt-4 were grown as described previously [34]. HeLa cells
were grown in DMEM (Dulbecco’s modified Eagle’s medium).
Where indicated, Molt-4, Jurkat, or HeLa cells were transfected
with either pSG5-Tax or empty vector using Lipofectamine Plus®
(Gibco, Invitrogen) according to manufacturer’s recommenda-
tions.
HPR (Sigma) was prepared as stock solutions in DMSO at 1 ×
102M and stored in amber tubes at 80 C. Before HPR ad-
dition, cells were cultured for 24 h at 1×105cells/ml in RPMI-
1640 medium containing 10 %(w/v) FBS (fetal bovine serum)
(Gibco, Invitrogen) and antibiotics. All experiments using HPR
were performed under yellow light (λ>500 nm) to prevent photo-
isomerization. PDMP (D,L-threo-1-phenyl-2-decanoylamino-3-
morpholino-1-propanol), purchased from Biomol (Plymouth
Meeting, PA, U.S.A.), was reconstituted in DMSO at a concen-
tration of 50 mM and was stored at 20 C. PDMP was added to
cells, 2 h before HPR treatment, to a final concentration ranging
from 10 µMto30µM. The final concentration of DMSO never
exceeded 0.1 %and this concentration showed no effect on the
proliferation of all tested cell lines (results not shown). The short-
chain cell-permeant ceramides C2-andC
6-ceramide or C2-and
Dh-C6(dihydro-C6-ceramide) (Biomol) were reconstituted in
100 %pure ethanol at a concentration of 40 mM, stored at
20 C and used at final concentrations ranging from 1 µMto
20 µM. Before ceramide and dihydroceramide treatments, cells
were cultured in RPMI-1640 medium containing 2 %(w/v) FBS.
FB1 (Biomol) was reconstituted in 1 ×PBS at a concentration
of 20 mM, stored at 20 C and added to cells at 50 µMfinal
concentration 1 h before HPR treatment.
Growth assays
Cell growth was assessed by cell counts using Trypan Blue dye
exclusion protocols and/or the use of the CellTiter 96®non-
radioactive cell proliferation assay kit (Promega Corp., Madison,
WI, U.S.A.) according to the manufacturer’s instructions. Cells
were grown in 96-well plates (Nunc, Naperville, IL, U.S.A.) and
at the initiation of cultures drugs were added at the concentrations
and time points indicated. Results are expressed as cell growth
relative to that in DMSO-treated controls and are derived from
the mean cell growth of quadruplicate wells. Results are represent-
ative of at least three independent experiments. Similar growth
trends were observed using the Trypan Blue dye exclusion and
the CellTiter 96®non-radioactive cell proliferation assays.
Ceramide measurement
Lipids were collected by the method of Bligh and Dyer [35].
Ceramide was measured with a modified DGK (diacylglycerol
kinase) assay using external ceramide standards as described
previously [36]. Briefly, 80 %of the lipid sample was dried under
N2. The dried lipid was solubilized in 20 µlofanoctylβ-D-
glucoside/dioleoyl phosphatidylglycerol micellar solution (7.5 %
octyl β-D-glucoside, 25 mM dioleoyl phosphatidylglycerol) by
several cycles of sonication in a bath sonicator followed by resting
at room temperature for 15–20 min. The reaction buffer was
prepared as a 2×solution, containing 100 mM imidazole HCl
(pH 6.6), 100 mM LiCl, 25 mM MgCl2and 2 mM EGTA. To the
lipid micelles, 50 µlof2×reaction buffer were added, 0.2 µlof
1M dithiothreitol, 5 µg of diglycerol kinase membranes and dilu-
tion buffer (10 mM imidazole, pH 6.6, 1 mM diethylenetriamine-
penta-acetic acid, pH 7) to a final volume of 90 µl. The reaction
was started by adding 10 µlof2.5mM[γ-32 P]ATP solution (spe-
cific activity of 75 000–200 000 c.p.m./nmol). The reaction was
allowed to proceed at 25C for 30 min. Bligh and Dyer lipid
extraction was performed and a 1.5 ml aliquot of the organic
phase was dried under N2. Lipids were then resuspended in
50 µl of methanol/chloroform (1:9, v/v) and 25 µl was spotted
on to a 20 cm silica gel TLC plate. Plates were developed using
chloroform/acetone/methanol/acetic acid/H2O (50:20:15:10:5, by
vol), air-dried and subjected to autoradiography. The radioactive
spots corresponding to phosphatidic acid and ceramide-phos-
phate, the phosphorylated products of diacylglycerol and cer-
amide respectively, were identified by comparison with known
standards. Spots were scraped into a scintillation vial containing
4 ml of scintillation fluid and counted on a scintillation counter.
Linear curves of phosphorylation were produced over a concen-
tration range of 0–960 pM of external standards (dioleoyl glycerol
and CIII-ceramide, Sigma). Ceramide levels were routinely
normalized to lipid phosphate levels. It is important to note that
under these conditions, there was a total conversion of ceramide
and diacylglycerol into their phosphorylated products, and there
was no change in the specific activity of the DGK enzyme.
De novo
ceramide synthesis
At initiation of treatment, [3H]palmitic acid (1 µCi/ml medium),
purchased from PerkinElmer (Boston, MA, U.S.A.) was added to
HPR-treated and control samples. Extracted lipids, dried under
N2, were resuspended in 1 ml of chloroform, 200 µlwasset
aside for the first phosphate measurement and 800 µl was used in
the base hydrolysis assay. In this assay, 200 µl of chloroform
and 800 µl of 0.125 M methanolic KOH were added to the
samples before incubation at 37 C for 75 min. The base was
neutralized by adding 100 µl of 1 M methanolic HCl to each
sample. A 2 ml volume of chloroform and 600 µl of distilled H2O
c
2005 Biochemical Society
Tax protein-induced defect in ceramide synthesis 233
were next added to the samples at room temperature for 30 min.
Subsequently, lipids were extracted by the method of Bligh and
Dyer [35], dried under N2and resuspended in 75 µl of chloroform.
A50µl sample of hydrolysed lipids was subjected to separation
for 1 h on a 20 cm silica gel TLC plate and 25 µlwaskeptfor
a second phosphate assay. The solvent system employed in
TLC separation was chloroform/methanol/2 M NH4OH (4:1:0.1,
by vol). The TLC plates were then sprayed lightly with En3hance®
(PerkinElmer) to enhance tritium readings. The ceramide spots
were visualized by iodine vapour mark. Radioactivity was also
visualized by autoradiography after 48 and 96 h at 80Candthe
[3H]ceramide spots were scraped into scintillation vials containing
4 ml of scintillation fluid and counted on a Packard scintillation
counter. Ceramide was quantified using external standards and
was normalized to phosphate counts.
Treatment with 17C6-ceramide
Hela cells, seeded at 2 ×106cells/ml, were transfected with
either pSG5-Tax or empty vector using Lipofectamine Plus®
(Gibco) according to the manufacturer’s recommendations. At
24 h post-transfection, cells were treated with D-erythro-17C6-
ceramide (generously provided by Dr Alicia Bielawska from the
Lipidomics Core at the Medical University of South Carolina,
Charleston, South Carolina, U.S.A.) for a further 24 h. This
unnatural ceramide undergoes deacylation and reacylation shortly
after cell entry where the generated 17-sphingosine backbone is
recycled and the activity of different ceramide synthases results in
the formation of several species of endogenous 17-ceramide. Cell
pellets were spiked with internal standards, lipids were extracted
and the unnatural 17-ceramide molecular species reflecting the
activity of ceramide synthases were analysed by MS. Data were
normalized to total protein levels.
Microsome preparation
Microsomes were prepared by sonication of cell pellets from CEM
and HuT-102 cells in 25 mM Tris, pH 7.4, 5 mM EDTA, 1 mM
PMSF and 1×complete protease inhibitor (Roche, Mannheim,
Germany)-containing buffer. The preparation was centrifuged at
1000 g, and the resulting supernatant was then ultracentrifuged
at 100 000 g. The pellet was resuspended by homogenization
in Hepes buffer (20 mM Hepes, 2 mM KCl and 2 mM MgCl2)
at pH 7.4. Protein concentration was determined using the DC
protein assay (Bio-Rad Laboratories, Hercules, CA, U.S.A.).
In vitro
ceramide synthase activity assay
Ceramide synthase activity was determined using microsome
fractions isolated from HPR-treated and control cells. A 50 µg
sample of microsomal proteins was incubated in 25 mM potas-
sium phosphate, pH 7.4, containing 50 µM palmitoyl-CoA and
15 µMD-erythro-17-sphingosine (a kind gift from Dr Alicia
Bielawska) delivered using BSA as vehicle. The incubation was
performed for 15 min at 37 C and was stopped by the addition of
the extraction solvent. Extracted lipids were used to measure the
levels of 17C16-ceramide by MS.
MS analysis of lipids
Samples were fortified with internal standards and lipids were
extracted with ethyl acetate/isopropanol/water (60:30:10 by vol),
evaporated to dryness and reconstituted in 100 µl of methanol.
Analysis was performed using electrospray ionization MS/MS
analysis on a Thermo Finnigan TSQ 7000 triple quadruple
mass spectrometer, operating in multiple-reactions-monitoring
positive-ionization mode, as described previously [37].
Statistical analysis
Three statistical tests were performed to validate the significance
of the observed results: the Dunnett test, the two-way analysis of
variance test, and the two-sample Students t-test.
RESULTS
HPR induces distinct ceramide responses in HTLV-I-negative and
HTLV-I-positive malignant T-cells
The effects of HPR treatment on ceramide levels was examined
in HTLV-I-negative cells. A time-dependent increase in ceramide
was observed in CEM (Figure 1A), Jurkat and Molt-4 cells (results
not shown) treated with HPR. This increase in ceramide levels
started after as little as 6 h in Jurkat cells and 12 h in CEM cells,
reaching a 2-fold increase after 18 h and increased to at least 3-
fold after 24 h in both cell lines. Ceramide generation preceded
the significant HPR-induced growth suppression and cell death
observed in these cell lines at the designated concentrations
(Figure 1B). A dose-dependent accumulation of ceramide was
noted in CEM cells treated for 24 h with increasing concentrations
of HPR, ranging from 0.1 µMto5µM (Figure 1C). In contrast,
no increase in ceramide levels was observed in HTLV-I-positive
HuT-102 (Figure 1D), C8166, and MT-2 cells (results not shown)
after 24 h of HPR treatment at 5 µM despite significant growth
suppression and cell death observed under these conditions [8].
Only late, modest increases in ceramide levels were observed in
HTLV-I-positive cells treated with 5 µMHPRfor48h(P<0.01,
t-test) (Figure 1D), after significant cell death had occurred [8].
Also, lower HPR doses had no effect (P<0.01, Dunnett test)
(Figure 1E).
HPR induces
de novo
synthesis of ceramide in HTLV-I-negative,
but not in HTLV-I-positive, malignant T-cells
Previous studies in different cell lines indicated that HPR in-
duces de novo synthesis of ceramide [30–32]. We examined the
role of de novo ceramide synthesis following HPR treatment
of HTLV-I-negative and HTLV-I-positive cell lines. CEM and
C8166 cells were treated with 1 and 5 µM HPR concentrations
respectively, based on their differential HPR sensitivity [8] at
indicated time points, and de novo synthesized ceramide was
measured by [3H]palmitate incorporation. HPR treatment caused
a time-dependent increase in [3H]ceramide, indicating de novo
synthesis in CEM cells, starting at 18 h post-treatment (P<0.01,
Dunnett test) (Figure 2). In contrast, C8166 cells showed no
change in [3H]ceramide after treatment using a 5-fold higher
HPR concentration (Figure 2). The activity of ceramide synthase
in CEM and HuT-102 cells in response to HPR treatment was
129.1 +
5.4 and 89.3 +
2.8 pmol of ceramide/min per mg of
protein respectively. There was no significant induction of this
activity when compared to the corresponding untreated cells.
These results suggest that de novo ceramide synthesis was im-
paired in HTLV-I-infected cells.
Inhibition of GCS (glucosylceramide synthase) increases
sensitivity to HPR-induced cell death in HTLV-I-positive cells
We next aimed to determine whether increasing the endogenous
levels of ceramide in HTLV-1-positive cells renders them more
sensitive to HPR. We therefore investigated the effect of PDMP, an
c
2005 Biochemical Society
234 N. Darwiche and others
Figure 1 HPR induces distinct ceramide responses in HTLV-I-negative and HTLV-I-positive malignant human T-cell lines
(A) Ceramide levels in the HTLV-I-negative CEM human T-cell line in response to HPR. CEM cells were seeded at a density of 3.5 ×105cells/ml and treated with 0.1% DMSO as a control or with
1µM HPR for the times indicated. Ceramide levels were determined using the DGK assay as described in the Experimental section and normalized to total cellular lipid phosphate levels. Data points
represent the mean +
range (
n
=2). Results are representative of three independent experiments. (B) Effects of HPR treatment on the growth of CEM human T-cell line. Cells were seeded at a density
of 2 ×105cells/ml and treated with 0.1 % DMSO as a control or with 1 µM HPR for the times indicated. Viable cell counts, calculated from triplicate wells by Trypan Blue dye exclusion, are expressed
as a percentage of controls and are representative of two independent experiments. (C) Dose–response to HPR treatment in CEM human T-cell line. Cells were seeded at a density of 3.5×105
cells/ml and treated with 0.1% DMSO as a control, or the indicated concentrations of HPR, for 24 h. Ceramide levels were determined as in (A). Data points represent the mean+
range (
n
=2).
Results are representative of three independent experiments. (D) HPR treatment induces late ceramide accumulation in HTLV-I-positive human T-cell line Hut-102. Cells were seeded at a density of
3.5 ×105cells/ml and treated with 0.1% DMSO as a control or 5 µM HPR for the times indicated. Ceramide levels were determined as in (A). Data points represent the mean+
range (
n
=2).
Results are representative of two independent experiments. (E) Dose–response to HPR treatment in HTLV-I-positive human T-cell line C8166. Cells were seeded at a density of 3.5 ×105cells/ml
and treated with the indicated concentrations of HPR for 48 h. Ceramide levels were determined as in (A). Data points represent the mean +
range (
n
=2). Results are representative of two independent
experiments.
Figure 2 HPR treatment induces
de novo
synthesis of ceramide in HTLV-I-
negative human T-cell lines only
CEM and C8166 cells were seeded at a density of 3.5 ×105cells/ml and treated with 0.1 % DMSO
as a control or with 1 and 5 µM HPR respectively.
De novo
ceramide levels were determined
using the [3H]palmitate incorporation method as described in the Experimental section and
normalized to total cellular lipid phosphate levels. Data points represent the mean +
range
(
n
=2) and are percentages of treated over control cells. Results are representative of two
independent experiments.
inhibitor of GCS [38] and thus an inducer of ceramide accumul-
ation, on HPR-induced cytotoxicity. At 2 h before HPR ex-
posure, cells were pretreated with PDMP at non-cytotoxic con-
centrations: 20 µM in C8166 and HuT-102 and 30 µMinMT2
cells. These concentrations resulted in a modest elevation of
cellular ceramide levels (Figure 3A, and results not shown).
The combination of HPR and PDMP in HTLV-I-positive cells
resulted in a reduction in cell viability (P<0.01, t-test) exceeding
their additive effects (Figure 3B). Treatment with 1 µMHPR
and 20 µM PDMP for 24 h led to a 12 %and 16 %synergistic
reduction in cell viability in C8166 and HuT-102 cells respectively
(P<0.01, t-test) (Figure 3B). Because of their lower sensitivity
to both PDMP and HPR [8], MT2 cells were treated with
3µMHPRand30µM PDMP for 24 h, resulting in a significant
21 %synergistic reduction in cell viability (P<0.01, t-test)
(results not shown). Only additive effects of HPR and PDMP
treatments were observed in the HTLV-I-negative cell line CEM,
although ceramide levels increased by approx. 100%over
control in response to PDMP (results not shown). These results
indicate that increasing endogenous ceramide levels by PDMP
partially restores the sensitivity of HTLV-I-positive cells to HPR
treatment.
Exogenous ceramides induce cell death in HTLV-I-negative and
HTLV-I-positive malignant T-cells
The results above suggested that HTLV-I-positive cells may have
a defect in ceramide synthesis in response to HPR and that
elevation of endogenous ceramide by PDMP renders them more
sensitive to HPR. Therefore it became important to determine
the sensitivity of both HTLV-I-positive and HTLV-I-negative
cells to exogenous ceramide as this might bypass the defect.
Addition of cell-permeant C2-andC
6-ceramide analogues triggers
many of the biological responses of agonists in cells [22]. We
treated the HTLV-I-negative Jurkat cells and the HTLV-I-positive
HuT-102, MT2 and C8166 cells with increasing concentrations
of C2-ceramide, ranging from 1 to 20 µM for up to 3 days. In
all cells tested, C2-ceramide caused a reduction in their viability
in a time- and dose-dependent manner (Table 1). A 20 µMC
2-
ceramide concentration completely killed the majority of tested
cells by 24 h. When the effects of the same concentration of
C2-ceramide in both cell type groups were compared, HTLV-I-
positive cells were slightly more resistant than HTLV-I-negative
cells. Additionally, cell treatment with Dh-C6, the immediate
inactive precursor of ceramide, ranging from 1 to 20 µMfor
up to 3 days, showed no effect on cell viability in any cell line,
underscoring the specificity of ceramide action (Table 1).
c
2005 Biochemical Society
Tax protein-induced defect in ceramide synthesis 235
Figure 3 PDMP synergizes with HPR treatment to induce cell death in
HTLV-I-positive human T-cell lines
(A) Ceramide levels in the HTLV-I-positive C8166 and HuT-102 cell lines in response to
combined HPR and PDMP treatment. Cells were seeded at a density of 3.5 ×105cells/ml and
were pretreated with 20 µM PDMP, an inhibitor of GCS. PDMP pretreatment was performed
2 h before treatment with 1 µM HPR. Data points represent the mean +
range (
n
=2). Results
are representative of two independent experiments. (B) Effect of combined HPR and PDMP
treatment on the growth of the HTLV-I-positive C8166 and HuT-102 cells. Cells were seeded
at a density of 2 ×105cells/ml and were p retreated with 20 µM PDMP, an inhibitor of GCS.
PDMP pretreatment was performed 2 h prior to treatment with 1 µM HPR. Viability was assayed
in 96-well plates at 24 h using the Cell Titer 96®non-radioactive cell proliferation kit. Data
points represent the mean +
S.D. (
n
=3). Results are expressed as percentage of control cells
(0.1% DMSO) and represent the mean of three independent experiments.
Recently, C6-ceramides were shown to induce the generation
of endogenous long-chain ceramides [39]. Our previous results
demonstrated that the HTLV-I-positive cells are not capable of
HPR-induced de novo ceramide synthesis and therefore may be
relatively less sensitive to C6-ceramides than the HTLV-I-negative
cells. We treated several HTLV-I-positive HuT-102, MT2 and
C8166 cells and HTLV-I-negative Jurkat and CEM cells with C6-
ceramide, ranging from 1 to 20 µM for up to 3 days. C6-ceramide
caused a time- and dose-dependent reduction in the growth of
all cells tested (Table 1). Viability of HTLV-I-positive-C8166
cells was measured using the CellTiter 96®non-radioactive cell
proliferation assay owing to excessive clumping of cells observed
in this cell line. In response to 20 µMC
2-ceramide, viability
decreased to 49 %,2%and 0 %of control at 24, 48 and 72 h
after treatment respectively. In response to the same concentration
of C6-ceramide, viability decreased to 69 %,23%and 18 %at
the same respective time points. Interestingly, when the effects
of the same concentration of C6-ceramide were compared in both
cell type groups, the HTLV-I-negative cells were found to be
significantly more sensitive to C6-ceramide than HTLV-I-positive
cells (Table 1). These results suggested, that in response to C6-
ceramide, HTLV-I-positive cells might generate lower levels of
endogenous ceramides.
Response to C6-ceramide in HTLV-I-positive cells is blunted,
resulting from a defect in endogenous ceramide synthesis
In order to verify whether ceramide synthesis in response to
exogenous C6-ceramide is affected in HTLV-I-positive cells, we
treated HuT-102, MT2 and C8166 cells, and the control CEM,
Jurkat and Molt-4 cells with 20 µMC
6-ceramide. Ceramide levels
were measured using the DGK assay, where the phosphorylated
products of C6-ceramide and endogenous ceramide can be identi-
fied by their distinct migration. C6-ceramide induced significantly
lower endogenous ceramide accumulation after 12 (results not
shown) and 18 h (Figure 4A) in all HTLV-I-positive cells tested,
compared with HTLV-I-negative cells. Elevated endogenous cer-
amide levels in HTLV-I-negative cells correlated with higher
levels of growth suppression and cell death in these cells after
12 (results not shown) and 18 h (Figure 4B) as compared with
HTLV-I-positive cells (C8166 viability decreased to 74%of
Table 1 Exogenous ceramides induce cell death in HTLV-I-negative and HTLV-I -positive malignant human T-cell lines
Effects of synthetic ceramides C2and C6on the growth of the HTLV-I-negative (Jurkat) and the HTLV-I-positive (HuT-102 and MT2) human T-cell lines. Cells were seeded in RPMI 1640 medium
containing 2 % (w/v) FBS in 24-well plates at a density of 2 ×105cells/ml and were treated with the indicated concentrations of C2-ceramide, C6-ceramide and Dh-C6. Viability was assayed using the
Trypan Blue dye exclusion method. Data points represent the mean viability (
n
=3). Results are expressed as percentage of control cells (0.1 % ethanol) and represent the mean of three independent
experiments.
Treatment time (h)
24 48 72
Cells Ceramide (µM) C2C6Dh-C6C2C6Dh-C6C2C6Dh-C6
JURKAT 1 µM 90 90 100 62 80 100 83 100 99
5µM 879599 386598 247096
10 µM 408298 116197 35595
20 µM009500950094
HuT-102 1 µM 92 94 100 67 65 100 56 100 100
5µM 556098 385698 299296
10 µM 315197 104296 07894
20 µM 8 42 95 0 35 94 0 52 94
MT2 10 µM 819397 798095 916495
15 µM 327096 206495 135895
20 µM 12 65 96 1 39 94 0 25 94
c
2005 Biochemical Society
236 N. Darwiche and others
Figure 4 Response to exogenous C6-ceramide in HTLV-I-positive human T-cell lines is blunted because of a defect in endogenous ceramide synthesis
(A) Endogenous ceramide levels in the HTLV-I-positive(HuT-102, MT2 and C8166) and the HTLV-I-negative(CEM, Jurkat and Molt-4) human T-cell lines in response to C6-ceramide. Cells were seeded
in RPMI-1640 medium containing 2 % FBS at a density of 3.5×105cell/ml. Cells were treated with 20 µMC
6-ceramide for 18 h. Endogenous ceramide levels were determined as in Figure 1(A). The
phosphorylated product of C6-ceramide is easily distinguishable from that of endogenous ceramide on the TLC plate (results not shown). Data points represent the mean +
range (
n
=2). Results are
representative of two independent experiments. (B) Effects of C6-ceramide on the growth of the HTLV-I-positive (HuT-102 and MT2) and the HTLV-I-negative (CEM, Jurkat and Molt-4) human T-cell
lines. Cells were seeded in RPMI-1640 medium containing 2 % FBS at a density of 2×105cells/ml. Cells were treated with 20 µMC
6-ceramide for 18 h. Viability was assayed using the Trypan Blue
dye exclusion method. Data points represent the mean +
S.D. (
n
=3). Results are expressed as percentage of control cells (0.1% ethanol) and represent the mean of two independent experiments.
(C) Endogenous ceramide levels in the HTLV-I-positive HuT-102 and the HTLV-I-negative Jurkat human T-cell lines in response to C6-ceramide and FB1treatment. Cells were seeded in RPMI-1640
medium containing 2 % FBS at a density of 3.5×105cell/ml. Cells were treated with 20 µMC
6-ceramide, 50 µMFB
1, or combined C6-ceramide and FB1for 18 h. Endogenous ceramide levels were
determined as in Figure 1(A). Data points represent the mean +
range (
n
=2). Results are representative of two independent experiments. (D) Endogenous ceramide levels in the HTLV-I-positive
HuT-102 and the HTLV-I-negative Jurkat human T-cell lines in response to exogenous C2-ceramide. Cells were seeded in RPMI-1640 medium containing 2 % FBS at a density of 3.5×105cell/ml.
Cells were treated with 20 µMC
2-ceramide for 18 h. Endogenous ceramide levels were determined as in Figure 1(A). Data points represent the mean +
range (
n
=2). Results are representative of
two independent experiments.
controls as measured using the CellTiter 96®non-radioactive cell
proliferation assay kit). To investigate the role of the de novo
ceramide synthetic pathway upon exogenous C6-ceramide ad-
dition, HuT-102 and Jurkat cells were pretreated with FB1,an
inhibitor of ceramide synthase. The addition of FB1almost com-
pletely abrogated the endogenous ceramide levels that were gene-
rated at 12 and 18 h after C6-ceramide addition in both cell lines
(Figure 4C). The viability of HPR-treated Jurkat and HuT-102
cells for up to 12 h increased by 35 %and 14 %respectively, fol-
lowing FB1pre-treatment, suggesting that endogenous ceramide
accumulation contributes significantly to HPR-induced cell death
of HTLV-I-negative cells.
The more potent growth suppressive effects of C2-ceramide
compared to C6-ceramide in both HTLV-I-positive and HTLV-I-
negative cells (Table 1) led us to examine whether the activity
of this analogue was less dependent on endogenous ceramide
generation. We treated HuT-102 and Jurkat cells with 20 µMC
2-
ceramide to evaluate its ability to induce endogenous ceramide
synthesis. Indeed, lower levels of endogenous ceramide levels
accumulated in the HTLV-I-positive cells after C2-ceramide ad-
ministration versus C6-ceramide (compare Figure 4D with Fig-
ure 4C), suggesting an explanation for the improved activity of this
analogue in HTLV-I-positive cells. Taken together, these results
clearly demonstrate that the HTLV-I-positive cells are defective
in endogenous ceramide synthesis, not only in response to HPR,
but also in response to exogenous ceramides.
Tax protein represses the ceramide response in
HTLV-I-positive cells
In order to test whether, in HTLV-I-infected cells, the oncoprotein
Tax is responsible for inhibiting ceramide accumulation, we
treated Tax-transfected Molt-4 cells with HPR (5 and 10 µM)
and measured ceramide levels. The levels of ceramide generated
were significantly lower in Tax-transfected cells in response to
both HPR concentrations after 14 and 18 h (P<0.05, t-test)
(Figure 5A). However, 5 µMHPRtreatmentfor20and38hinthe
HTLV-I-negative CEM cells and HTLV-I-positive HuT-102 cells
did not significantly stimulate ceramide synthase activity (results
not shown). Furthermore, HeLa cells transiently transfected
with Tax protein had a significantly reduced accumulation of
endogenous ceramide after 18 h (Figure 5B) in response to
c
2005 Biochemical Society
Tax protein-induced defect in ceramide synthesis 237
Figure 5 Role of Tax protein in suppressing the ceramide response in the
HTLV-I-positive human T-cell lines
(A) Reduced ceramide levels in HPR-treated Tax-transfected Molt-4 cells. Molt-4 cells were
transiently transfected with pSG5-Taxor control vector and treated after 24 h with HPR (5 µMand
10 µM) for the times indicated. Ceramide levels were determined as in Figure 1(A). Data points
represent the mean +
range (
n
=2). Results are representative of two independent experiments.
(B) Effects of exogenous C6-ceramide on endogenous ceramide levels in Tax-transfected HeLa
cells. pSG5-Tax or control cells were transfected for 24 h and then were treated with 20 µM
C6-ceramide in DMEM medium containing 2 % FBS for 18 h. Endogenous ceramide levels
were determined as in Figure 1(A). Data points represent the mean +
range (
n
=2). Results
are representative of two independent experiments. (C) Ceramide synthase activity is reduced
in Tax-transfected Hela cells. pSG5-Tax or control cells were transfected for 24 h and then were
treated with 5 µM 17C6-ceramide in DMEM medium containing 2 % FBS for 24 h. Endogenous
17-ceramide species levels were determined as described in the Experimental section and
normalized to total cellular protein levels. Data points represent the mean +
range (
n
=2).
C6-ceramide treatment (P<0.01, t-test). Moreover, we utilized
the synthetic ceramide 17C6-ceramide, that has an unnatural
17-sphingosine backbone, to indirectly determine the activity of
ceramide synthase(s). Measurement of the unnatural endogenous
17-ceramides that result from the deacylation/reacylation reaction
(Figure 5C) after 24 h showed that in Tax-transfected cells, the
production of most endogenous 17-ceramides was significantly
inhibited (P<0.01 for all except 17C24-ceramide, t-test). These
experiments when combined with the results above indicate that
Tax is responsible for suppression of de novo ceramide synthesis,
probably by inhibiting ceramide synthase.
DISCUSSION
We have determined that HPR induces distinct ceramide res-
ponses in HTLV-I-negative and HTLV-I-positive malignant T-
cells. This may be partly explained by an impaired de novo
ceramide synthetic pathway in the HTLV-I-positive cells owing
to expression of Tax protein. De novo synthesis of ceramide is
dependent on the activity of several enzymes, including several
ceramide synthases. Recently, ceramide synthase genes, identified
as Lass (longevity assurance-like) genes, were described but have
not yet been well characterized [40]. The results from this study
indicate that the activity of ceramide synthase(s) is suppressed in
HTLV-I-positive cells.
The HTLV-I-encoded oncoprotein, Tax, plays a key role in
the transformation of infected cells and their resistance to chemo-
therapy [8,41]. In the current study, it is shown that Tax expression
in cells not infected with HTLV-I is sufficient to inhibit the
generation of ceramide in response to HPR or C6-ceramide and to
suppress ceramide synthase activity as measured by the ability of
transfected cells to convert the 17C6-ceramide into endogenous
17-ceramides. A major biological function of Tax protein appears
to be its ability to induce a constitutive high level of activity
of the transcription factor NF-κB (nuclear factor-κB) [42].
Recent evidence suggests that the Tax oncoprotein represses
the transcriptional activity of the tumour suppressor protein p53
through the activation of the NF-κB pathway, but independent
of the CREB/ATF pathway, which is also disrupted by Tax [43].
Interestingly, p53 wasfound to regulate the generation of ceramide
in response to chemotherapeutic agents or γ-irradiation [44].
Therefore one future aim will be to examine whether NF-κBand
p53 mediate the inhibitory effects of Tax on de novo synthesis of
ceramide at the genetic level.
Cell permeant ceramide analogues C2 and C6 were shown to
reproduce the growth suppressive effects of chemotherapeutic
agents that induce ceramide accumulation [27]. These short-
chain ceramides have also been employed to support evidence for
apoptosis induction obtained after elevation of cellular ceramide
content in response to several inducers of ceramide accumulation
such as TNF-α, Fas ligand, interleukin-1 and other effectors
[45–47]. C6-ceramide was recently shown to induce de novo
ceramide synthesis in the A549 human lung adenocarcinoma
cell line where its sphingosine backbone becomes recycled after
deacylation/reacylation into endogenous long-chain ceramides
[39]. This revealed that the biological functioning of exogenous
ceramides was in part dependent on ceramide synthase activity
to produce endogenous ceramides. In the current study C2-as
well as C6-ceramides induced de novo ceramide synthesis in
both HTLV-I-positive and HTLV-I-negative cells. However, this
was markedly lower in HTLV-I-positive cells, and in both cell
types, was lower in response to the C2-ceramide analogue. Both
analogues induced growth suppression in both cell types in a time-
and dose-dependent manner. The lower sensitivity of HTLV-I-
positive cells to exogenous ceramides provided further support
for the suggestion that ceramide synthase is inhibited in these
cells. It was also noted, that under the same conditions, C2-
ceramide was more effective compared with C6-ceramide in
producing growth suppression. These differences could be due
to better uptake, slower metabolism, or an enhanced direct effect
for this analogue, as opposed to dependence on the generation of
endogenous ceramide.
c
2005 Biochemical Society
238 N. Darwiche and others
Several studies indicate that the co-administration of modul-
ators of ceramide metabolism and chemotherapeutic agents in-
creases the level of ceramide and further enhances cytotoxicity.
PDMP and related compounds are potent inhibitors of GCS
and consequently of ceramide clearance. These inhibitors have
been shown to sensitize a variety of drug-resistant cell lines to
chemotherapeutic reagents and cause their preferential killing
by such reagents [26,48,49]. Furthermore, D,L-threo-(1-phenyl-
2-hexadecanoylamino)-3-morpholino-1-propanol administration
produces a synergistic increase in cytotoxicity with HPR in
solid tumour cell lines [31] and in HPR-resistant PC-3 prostate
cancer cells [33]. In the present study HTLV-I-positive cells
were found to have a defect in ceramide synthesis and relative
resistance to HPR as compared with HTLV-I-negative cells. Re-
storation of elevated ceramide levels by PDMP during HPR treat-
ment led to synergistic cytotoxicity in HTLV-I-positive cell lines,
thus sensitizing these cells to HPR concentrations that are lower
than what is usually required to induce significant growth sup-
pression. This, together with results using exogenous ceramides,
indicated that HTLV-I-positive cells were sensitive to cera-
mide-induced cell killing once sufficient levels were generated.
Deregulation of ceramide metabolism may act as an indicator
of chemoresistance in tumour cells and could be a target in
cancer therapy [28,50]. The present study demonstrates that viral
oncogenes may target the ceramide pathway. In the case of Tax,
this occurs by inhibition of ceramide synthase activity and may
explain the resistance of HTLV-1-positive cells to treatment.
Therapeutic agents with the ability to bypass this defect are more
likely to be successful.
This work was supported by grants from the Lebanese National Council for Scientific
Research and from the American University of Beirut, University Research Board (N. D.)
and by the National Science Foundation (U.S.A.) Grant No. 0090859 (G.S.D.). We thank
Dr Jacek Bielawski for performing MS analysis of ceramides at the Lipidomics Core
Laboratory at the Medical University of South Carolina (M.U.S.C.) and Dr Alicia Bielawska
(M.U.S.C.) for providing the 17C6-ceramide and 17-sphingosine.
REFERENCES
1 Delia, D., Aiello, A., Lombardi, L., Pelicci, P. G., Grignani, F., Grignani, F., Formelli, F.,
Menard, S., Costa, A. and Veronesi, U. (1993) N-(4-hydroxyphenyl)retinamide induces
apoptosis of malignant hemopoietic cell lines including those unresponsive to retinoic
acid. Cancer Res. 53, 6036–6041
2 Wu, J. M., DiPietrantonio, A. M. and Hsieh, T. C. (2001) Mechanism of fenretinide
(4-HPR)-induced cell death. Apoptosis 6, 377–388
3 Kalemkerian, G. P., Slusher, R., Ramalingam, S., Gadgeel, S. and Mabry, M. (1995)
Growth inhibition and induction of apoptosis by fenretinide in small-cell lung cancer cell
lines. J. Natl. Cancer Inst. 87, 1674–1680
4 Maurer, B. J., Metelitsa, L. S., Seeger, R. C., Cabot, M. C. and Reynolds, C. P. (1999)
Increase of ceramide and induction of mixed apoptosis/necrosis by N-(4-
hydroxyphenyl)-retinamide in neuroblastoma cell lines. J. Natl. Cancer Inst. 91,
1138–1146
5 Oridate, N., Suzuki, S., Higuchi, M., Mitchell, M. F., Hong, W. K. and Lotan, R. (1997)
Involvement of reactive oxygen species in N-(4-hydroxyphenyl)retinamide-induced
apoptosis in cervical carcinoma cells. J. Natl. Cancer Inst. 89, 1191–1198
6 Chen, Y. R., Zhou, G. and Tan, T. H. (1999) c-Jun N-terminal kinase mediates apoptotic
signaling induced by N-(4-hydroxyphenyl)retinamide. Mol. Pharmacol. 56,
1271–1279
7 Erdreich-Epstein, A., Tran, L. B., Bowman, N. N., Wang, H., Cabot, M. C., Durden, D. L.,
Vlckova, J., Reynolds, C. P., Stins, M. F., Groshen, S. and Millard, M. (2002) Ceramide
signaling in fenretinide-induced endothelial cell apoptosis. J. Biol. Chem. 277,
49531–49537
8 Darwiche, N., Hatoum, A., Dbaibo, G., Kadara, H., Nasr, R., Abou-Lteif, G., Bazzi, R.,
Hermine, O., de The, H. and Bazarbachi, A. (2004) N-(4-hydroxyphenyl)retinamide
induces growth arrest and apoptosis in HTLV-I-transformed cells. Leukemia 18,
607–615
9 Hannun, Y. A. (1996) Functions of ceramide in coordinating cellular responses to stress.
Science (Washington D.C.) 274, 1855–1859
10 Andrieu-Abadie, N., Gouaze, V., Salvayre, R. and Levade, T. (2001) Ceramide in apoptosis
signaling: relationship with oxidative stress. Free Radicals Biol. Med. 31, 717–728
11 Perry, D. K. (2000) The role of de novo ceramide synthesis in chemotherapy-induced
apoptosis. Ann. N.Y. Acad. Sci. 905, 91–96
12 Okazaki, T., Bielawska, A., Bell, R. M. and Hannun, Y. A. (1990) Role of ceramide as a lipid
mediator of 1α,25-dihydroxyvitamin D3-induced HL-60 cell differentiation. J. Biol. Chem.
265, 15823–15831
13 Dbaibo, G. S., Pushkareva, M. Y., Jayadev, S., Schwarz, J. K., Horowitz, J. M.,
Obeid, L. M. and Hannun, Y. A. (1995) Retinoblastoma gene product as a downstream
target for a ceramide-dependent pathway of growth arrest. Proc. Natl. Acad. Sci. U.S.A.
92, 1347–1351
14 Lee, J. Y., Leonhardt, L. G. and Obeid, L. M. (1998) Cell-cycle-dependent changes in
ceramide levels preceding retinoblastoma protein dephosphorylation in G2/M.
Biochem. J. 334, 457–461
15 Ogretmen, B., Kraveka, J. M., Schady, D., Usta, J., Hannun, Y. A. and Obeid, L. M. (2001)
Molecular mechanisms of ceramide-mediated telomerase inhibition in the A549 human
lung adenocarcinoma cell line. J. Biol. Chem. 276, 32506–32514
16 Pettus, B. J., Chalfant, C. E. and Hannun, Y. A. (2002) Ceramide in apoptosis: an overview
and current perspectives. Biochim. Biophys. Acta. 1585, 114–125
17 Gulbins, E. (2003) Regulation of death receptor signaling and apoptosis by ceramide.
Pharmacol. Res. 47, 393–399
18 Garc´
ıa-Ruiz, C., Colell, A., Mar´
ı, M., Morales, A. and Fern´
andez-Checa, J. C. (1997)
Direct effect of ceramide on the mitochondrial electron transport chain leads to generation
of reactive oxygen species. Role of mitochondrial glutathione. J. Biol. Chem. 272,
11369–11377
19 Ruvolo, P. P. (2003) Intracellular signal transduction pathways activated by ceramide and
its metabolites. Pharmacol. Res. 47, 383–392
20 Zhao, S., Yang, Y. N. and Song, J. G. (2004) Ceramide induces caspase-dependent and
-independent apoptosis in A-431 cells. J. Cell Physiol. 199, 47–56
21 Hannun, Y. A. and Luberto, C. (2000) Ceramide in the eukaryotic stress response.
Trends Cell Biol. 10, 73–80
22 Cai, Z., Bettaieb, A., Mahdani, N. E., Legres, L. G., Stancou, R., Masliah, J. and
Chouaib, S. (1997) Alteration of the sphingomyelin/ceramide pathway is associated with
resistance of human breast carcinoma MCF7 cells to tumor necrosis factor-α-mediated
cytotoxicity. J. Biol. Chem. 272, 6918–6926
23 Chmura, S. J., Nodzenski, E., Beckett, M. A., Kufe, D. W., Quintans, J. and
Weichselbaum, R. R. (1997) Loss of ceramide production confers resistance to
radiation-induced apoptosis. Cancer Res. 57, 1270–1275
24 Michael, J. M., Lavin, M. F. and Watters, D. J. (1997) Resistance to radiation-induced
apoptosis in Burkitt’s lymphoma cells is associated with defective ceramide signaling.
Cancer Res. 57, 3600–3605
25 Chmura, S. J., Mauceri, H. J., Advani, S., Heimann, R., Beckett, M. A., Nodzenski, E.,
Quintans, J., Kufe, D. W. and Weichselbaum, R. R. (1997)Decreasing the apoptotic
threshold of tumor cells through protein kinase C inhibition and sphingomyelinase
activation increases tumor killing by ionizing radiation. Cancer Res. 57, 4340–4347
26 Lavie, Y., Cao, H. t., Volner, A., Lucci, A., Han, T. Y., Geffen, V., Giuliano, A. E. and
Cabot, M. C. (1997) Agents that reverse multidrug resistance, tamoxifen, verapamil, and
cyclosporin A, block glycosphingolipid metabolism by inhibiting ceramide glycosylation
in human cancer cells. J. Biol. Chem. 272, 1682–1687
27 Bielawska, A., Crane, H. M., Liotta, D., Obeid, L. M. and Hannun, Y. A. (1993) Selectivity
of ceramide-mediated biology: lack of activity of erythro-dihydroceramide. J. Biol. Chem.
268, 26226–26232
28 Reynolds, C. P., Maurer, B. J. and Kolesnick, R. N. (2004) Ceramide synthesis and
metabolism as a target for cancer therapy. Cancer Lett. 206, 169–180
29 DiPietrantonio, A. M., Hsieh, T. C., Olson, S. C. and Wu, J. M. (1998) Regulation of G1/S
transition and induction of apoptosis in HL-60 leukemia cells by fenretinide (4HPR).
Int. J. Cancer 78, 53–61
30 Wang, H., Maurer, B. J., Reynolds, C. P. and Cabot, M. C. (2001)
N-(4-hydroxyphenyl)retinamide elevates ceramide in neuroblastoma cell lines by
coordinate activation of serine palmitoyltransferase and ceramide synthase. Cancer Res.
61, 5102–5105
31 Maurer, B. J., Melton, L., Billups, C., Cabot, M. C. and Reynolds, C. P. (2000) Synergistic
cytotoxicity in solid tumor cell lines between N-(4-hydroxyphenyl)retinamide and
modulators of ceramide metabolism. J. Natl. Cancer Inst. 92, 1897–1909
32 O’Donnell, P. H., Guo, W. X., Reynolds, C. P. and Maurer, B. J. (2002)
N-(4-hydroxyphenyl)retinamide increases ceramide and is cytotoxic to acute
lymphoblastic leukemia cell lines, but not to non-malignant lymphocytes. Leukemia 16,
902–910
33 Wang, H., Charles, A. G., Frankel, A. J. and Cabot, M. C. (2003) Increasing intracellular
ceramide: an approach that enhances the cytotoxic response in prostate cancer cells.
Urology 61, 1047–1052
c
2005 Biochemical Society
Tax protein-induced defect in ceramide synthesis 239
34 Darwiche, N., El-Sabban, M., Bazzi, R., Nasr, R., Al-Hashimi, S., Hermine, O., de The, H.
and Bazarbachi, A. (2001) Retinoic acid dramatically enhances the arsenic
trioxide-induced cell cycle arrest and apoptosis in retinoic acid receptor alpha-positive
human T-cell lymphotropic virus type-I-transformed cells. Hematol. J. 2, 127–135
35 Bligh, E. G. and Dyer, W. J. (1959) A rapid method of total lipid extraction and
purification. Can. J. Bioch. Phys. 37, 911–917
36 Bielawska, A., Perry, D. K. and Hannun, Y. A. (2001) Determination of ceramides and
diglycerides by the diglyceride kinase assay. Anal. Biochem. 298, 141–150
37 Pettus, B. J., Bielawski, J., Porcelli, A. M., Reames, D. L., Johnson, K. R., Morrow, J.,
Chalfant, C. E., Obeid, L. M. and Hannun, Y. A. (2003) The sphingosine kinase
1/sphingosine-1-phosphate pathway mediates COX-2 induction and PGE2 production in
response to TNF-alpha. F.A.S.E.B. J. 17, 1411–1421
38 Radin, N. S. (1994) Rationales for cancer chemotherapy with PDMP, a specific inhibitor of
glucosylceramide synthase. Mol. Chem. Neuropathol. 21, 111–127
39 Ogretmen, B., Pettus, B. J., Rossi, M. J., Wood, R., Usta, J., Szulc, Z., Bielawska, A.,
Obeid, L. M. and Hannun, Y. A. (2002) Biochemical mechanisms of the generation of
endogenous long chain ceramide in response to exogenous short chain ceramide in the
A549 human lung adenocarcinoma cell line. Role for endogenous ceramide in mediating
the action of exogenous ceramide. J. Biol. Chem. 277, 12960–12969
40 Riebeling, C., Allegood, J., Wang, E., Merrill, A. Jr and Futerman, A. (2003) Two
mammalian longevity assurance gene (LAG1) family members,
trh1
and
trh4
, regulate
dihydroceramide synthesis using different fatty Acyl-CoA donors. J. Biol. Chem. 278,
43452–43459
41 Gatza, M. L., Watt, J. C. and Marriott, S. J. (2003) Cellular transformation by the HTLV-I
Tax protein, a jack-of-all-trades. Oncogene 22, 5141–5149
42 Pise-Masison, C., Mahieux, R., Radonovich, M., Jiang, H. and Brady, J. (2001) Human
T-lymphotropic virus type I Tax protein utilizes distinct pathways for p53 inhibition that
are cell type-dependent. J. Biol. Chem. 276, 200–205
43 Pise-Masison, C. A., Mahieux, R., Jiang, H., Ashcroft, M., Radonovich, M., Duvall, J.,
Guillerm, C. and Brady, J. N. (2000) Inactivation of p53 by Human T-Cell -lymphotropic
virus type 1 Tax requires activation of the NF-κB pathway and is dependent on p53
phosphorylation. Mol. Cell. Biol. 20, 3377–3386
44 Dbaibo, G. S., Pushkareva, M. Y., Rachid, R. A., Alter, N., Smyth, M. J., Obeid, L. M. and
Hannun, Y. A. (1998) p53-dependent ceramide response to genotoxic stress.
J. Clin. Invest. 102, 329–339
45 Mathias, S., Younes, A., Kan, C.-C., Orlow, I., Joseph, C. and Kolesnick, R. N. (1993)
Activation of the sphingomyelin signaling pathway in intact EL4 cells and in a cell-free
system by IL-1β. Science (Washington D.C.) 259, 519–522
46 Cifone, M. G., De Maria, R., Roncaioli, P., Rippo, M. R., Azuma, M., Lanier, L. L.,
Santoni, A. and Testi, R. (1994) Apoptotic signaling through CD95 (Fas/Apo-1) activates
an acidic sphingomyelinase. J. Exp. Med. 180, 1547–1552
47 Tepper, C. G., Jayadev, S., Liu, B., Bielawska, A., Wolff, R., Yonehara, S.,
Hannun, Y. A. and Seldin, M. F. (1995) Role of ceramide as an endogenous
mediator of Fas-induced cytotoxicity. Proc. Natl. Acad. Sci. U.S.A. 92,
8443–8447
48 Nicholson, K. M., Quinn, D. M., Kellett, G. L. and Warr, J. R. (1999) Preferential killing of
multidrug-resistant KB cells by inhibitors of glucosylceramide synthase. Br. J. Cancer 81,
423–430
49 Spinedi, A., Bartolomeo, S. D. and Piacentini, M. (1998) Apoptosis induced by
N-hexanoylsphingosine in CHP-100 cells associates with accumulation of endogenous
ceramide and is potentiated by inhibition of glucocerebroside synthesis. Cell Death Differ.
5, 785–791
50 Batra, S., Reynolds, C. P. and Maurer, B. J. (2004) Fenretinide cytotoxicity for Ewing’s
sarcoma and primitive neuroectodermal tumor cell lines is decreased by
hypoxia and synergistically enhanced by ceramide modulators. Cancer Res. 64,
5415–5424
Received 8 April 2005/29 July 2005; accepted 9 August 2005
Published as BJ Immediate Publication 9 August 2005, doi:10.1042/BJ20050578
c
2005 Biochemical Society
... 1, Cer can be generated by 3 different pathways, i.e., de novo synthesis, sphingomyelin hydrolysis , and the recently described salvage pathway (Hannun and Obeid 2008; Kitatani et al. 2008). Numerous studies have implicated Cer in 4-HPR-mediated cytotoxicity (Maurer et al. 1999; Rehman et al. 2004; Darwiche et al. 2005; Hail et al. 2006; Morale s et al. 2007; Jiang et al. 2011 ) but application of more advanced liquid chromatography – mass spectrometry (LC–MS) techniques has recently revealed that 4- HPR induces an increase in dihydroceramide (dhCer) rather than Cer, at least in some tumor models (Kraveka et al. 2007; Wang et al. 2008; Valsecchi et al. 2010). The aim of this work was to establish the mechanistic relationship among the 4-HPR-induced oxidative stress, changes in SLs, and cell death. ...
... Therefore, the data support the potential application of fenretinide as a chemotherapeutic agent. SLs (especially Cer and gangliosides) have been largely linked to 4-HPR-driven cytotoxicity (Maurer et al. 1999; Wu et al. 2001; Lovat et al. 2004; Rehman et al. 2004; Darwiche et al. 2005; Hail et al. 2006; Morales et al. 2007). Previous studies based on thin layer chromatography pointed out an increase in endogenous Cer as a major response in 4-HPR- mediated cell death (Maurer et al. 1999; Chen et al. 2003; Rehman et al. 2004; Hail et al. 2006; Morales et al. 2007; Tiwari et al. 2008). ...
Article
Full-text available
4-(Hydroxyphenyl)retinamide (4-HPR) is a synthetic retinoid with a strong apoptotic effect towards different cancer cell lines in vitro, and it is currently tested in clinical trials. Increases of reactive oxygen species (ROS) and modulation of endogenous sphingolipid levels are well-described events observed upon 4-HPR treatment, but there is still a lack of understanding of their relationship and their contribution to cell death. LC-MS analysis of sphingolipids revealed that in human leukemia CCRF-CEM and Jurkat cells, 4-HPR induced dihydroceramide but not ceramide accumulation even at sublethal concentrations. Myriocin prevented the 4-HPR-induced dihydroceramide accumulation, but it did not prevent the loss of viability and increase of intracellular ROS production. On the other hand, ascorbic acid, Trolox, and vitamin E reversed 4-HPR effects on cell death but not dihydroceramide accumulation. NDGA, described as a lipoxygenase inhibitor, exerted a significantly higher antioxidant activity than vitamin E and abrogated 4-HPR-mediated ROS. It did not however rescue cellular viability. Taken together, this study demonstrates that early changes observed upon 4-HPR treatment, i.e., sphingolipid modulation and ROS production, are mechanistically independent events. Furthermore, the results indicate that 4-HPR-driven cell death may occur even in the absence of dihydroceramide or ROS accumulation. These observations should be taken into account for an improved design of drug combinations.
Article
Full-text available
Annual influenza outbreaks are associated with significant morbidity and mortality worldwide despite the availability of seasonal vaccines. Influenza pathogenesis depends on the manipulation of host cell signaling to promote virus replication. Ceramide is a sphingosine-derived lipid that regulates diverse cellular processes. Studies highlighted the differential role of ceramide de novo biosynthesis on the propagation of various viruses. Whether ceramide plays, a role in influenza virus replication is not known. In this study, we assessed the potential interplay between the influenza A (IAV) and ceramide biosynthesis pathways. Accumulation of ceramide in human lung epithelial cells infected with influenza A/H1N1 virus strains was evaluated using thin layer chromatography and/or confocal microscopy. Virus replication was assessed upon the regulation of the de novo ceramide biosynthesis pathway. A significant increase in ceramide accumulation was observed in cells infected with IAV in a dose- and time-dependent manner. Inoculating the cells with UV-inactivated IAV or treating them with sialidase prior to virus inoculation did not result in ceramide accumulation in the cells suggesting that the induction of ceramide required an active virus replication. Inhibiting de novo ceramide significantly decreased ceramide accumulation and enhanced virus replication. The addition of exogenous C 6 -ceramide prior to infection mediated an increase in cellular ceramide levels and significantly attenuated IAV replication and reduced viral titers (≈1log10 pfu/ml unit) Therefore, our data demonstrate that ceramide accumulation through de novo biosynthesis pathway plays a protective and antiviral role against IAV infection. These findings propose new avenues for development of antiviral molecules and strategies. Importance: Understanding the effect of sphingolipid metabolism on viral pathogenesis provide important insights into the development of therapeutic strategies against microbial infections. In this study, we demonstrate a critical role of ceramide during influenza A virus infection. We demonstrate that ceramide produced through de novo biosynthesis possess an antiviral role. These observations unlock new opportunities for the development of novel antiviral therapies against influenza.
Article
Full-text available
Early in apoptosis, ceramide levels rise and the mitochondrial outer membrane becomes permeable to small proteins. The self-assembly of ceramide to form channels could be the means by which intermembrane space proteins are released to induce apoptosis. Dihydroceramide desaturase converts dihydroceramide to ceramide. This conversion may be removing an inhibitor as well as generating a pro-apoptotic agent. We report that both long and short chain dihydroceramides inhibit ceramide channel formation in mitochondria. One tenth as much dihydroceramide was sufficient to inhibit the permeabilization of the outer membrane by about 95% (C2) and 51% (C16). Similar quantities inhibited the release of carboxyfluorescein from liposomes indicating that other mitochondrial components are not necessary for the inhibition. The apoptogenic activity of ceramide may thus depend on the ceramide to dihydroceramide ratio resulting in a more abrupt transition from the normal to the apoptotic state when the de novo pathway is used in mitochondria.
Article
Full-text available
N-(4-hydroxyphenyl)retinamide (HPR) is a synthetic retinoid that inhibits growth of many human tumor cells, including those resistant to natural retinoids. HPR is an effective chemopreventive agent for prostate, cervix, breast, bladder, skin and lung cancers, and has shown promise for the treatment of neuroblastomas. We have previously shown that HPR inhibits proliferation and induces apoptosis of human T-cell lymphotropic virus type I (HTLV-I)-associated adult T-cell leukemia (ATL) and HTLV-I-negative malignant T cells, whereas no effect is observed on normal lymphocytes. In this report, we identified HPR-induced reactive oxygen species (ROS) generation as the key mediator of cell cycle arrest and apoptosis of malignant T cells. HPR treatment of HTLV-I-negative malignant T cells was associated with a rapid and progressive ROS accumulation. Pre-treatment with the antioxidants vitamin C and dithiothreitol inhibited ROS generation, prevented HPR-induced ceramide accumulation, cell cycle arrest, cytochrome c release, caspase-activation and apoptosis. Therefore, anti-oxidants protected malignant T cells from HPR-induced growth inhibition. The expression of the HTLV-I oncoprotein Tax abrogated HPR-induced ROS accumulation in HTLV-I-infected cells, which explains their lower sensitivity to HPR. Defining the mechanism of free radical induction by HPR may support a potential therapeutic role for this synthetic retinoid in ATL and HTLV-I-negative T-cell lymphomas.
Article
Full-text available
Arsenic trioxide (ATO) is an effective treatment for acute promyelocytic leukemia (APL) and potentially for human T-cell leukemia virus type I (HTLV-I) associated adult T-cell leukemia/lymphoma (ATL). Many cytotoxic drugs induce apoptosis through the generation and accumulation of the sphingolipid breakdown product, ceramide, a coordinator of the cellular response to stress. We, therefore, investigated the contribution of ceramide to the mechanism of action of ATO in APL and ATL. A human APL-derived cell line (NB4), various ATL-derived lines and an HTLV-I-negative malignant T-cell line were cultured and treated with ATO. Growth and apoptosis assays were conducted. Measurements were made of ceramide, diacylglycerol, sphingomyelinase activity, sphingomyelin mass, glucosylceramide synthase activity and the de novo ceramide synthesis. Treatment of APL and ATL-derived cells with a clinically achievable concentration of ATO induced accumulation of cytotoxic levels of ceramide. The effects of ATO on ceramide levels in APL cells were more potent than those of all-trans retinoic acid (ATRA). ATO downregulated neutral sphingomyelinase activity. In contrast to the effect of ATRA, ATO-induced ceramide accumulation was not due to induction of acidic sphingomyelinase, but rather resulted from both de novo ceramide synthesis and inhibition of glucosylceramide synthase activity. Interestingly, the effects of ATO on de novo ceramide synthesis were similar in APL and ATL-derived cells despite the defective pathway in ATL cells. These results indicate that ATO-induced ceramide accumulation may represent a general mediator of the effects of ATO, which paves the way for new therapeutic interventions that target the metabolic pathway of this important sphingolipid secondary messenger.
Article
Full-text available
The treatment of HL-60 myelocytic leukemia cells with 1 alpha,25-dihydroxyvitamin D3 (1,25-(OH)2D3) resulted in the activation of a neutral sphingomyelinase and in sphingomyelin turnover (Okazaki, T., Bell, R., and Hannun, Y. (1989) J. Biol. Chem. 264, 19076-19080). In this paper, the effects of 1,25-(OH)2D3 on the product of sphingomyelin hydrolysis, ceramide, and the possible function of ceramide as a lipid mediator of the effects of 1,25-(OH)2D3 on HL-60 cell differentiation were investigated. Treatment of HL-60 cells with 1,25-(OH)2D3 resulted in a time- and dose-dependent increase in ceramide mass levels. Ceramide levels peaked at 2 h following treatment of HL-60 cells with 100 nM 1,25-(OH)2D3 with an increase of 41% over base line. The mass of generated ceramide (13 +/- 2 pmol/nmol of phospholipid) agreed with the mass of hydrolyzed sphingomyelin (17 +/- 4 pmol/nmol of phospholipid). Cell-permeable ceramides with shorter N-acyl chains induced HL-60 cell differentiation at subthreshold concentrations of 1,25-(OH)2D3. Higher concentrations of cell-permeable ceramides potently induced HL-60 cell differentiation independent of 1,25-(OH)2D3. A 2-h exposure of HL-60 cells to N-acetyl-sphingosine was sufficient to cause differentiation. Morphologically, N-acetylsphingosine caused a similar monocytic differentiation of HL-60 cells as did 1,25-(OH)2D3. Exogenous ceramide was further metabolized to sphingomyelin and other sphingolipids, but no conversion to sphingosine was detected. Moreover, sphingosine and its analogs failed to affect monocytic differentiation of HL-60 cells in response to subthreshold 1,25-(OH)2D3, indicating that the effect of ceramide was independent of sphingosine generation. These studies demonstrate that ceramide is a lipid mediator that may transduce the action of 1,25-(OH)2D3 on HL-60 cell differentiation.
Article
Full-text available
Intracellular pathways leading from membrane receptor engagement to apoptotic cell death are still poorly characterized. We investigated the intracellular signaling generated after cross-linking of CD95 (Fas/Apo-1 antigen), a broadly expressed cell surface receptor whose engagement results in triggering of cellular apoptotic programs. DX2, a new functional anti-CD95 monoclonal antibody was produced by immunizing mice with human CD95-transfected L cells. Crosslinking of CD95 with DX2 resulted in the activation of a sphingomyelinase (SMase) in promyelocytic U937 cells, as well as in other human tumor cell lines and in CD95-transfected murine cells, as demonstrated by induction of in vivo sphingomyelin (SM) hydrolysis and generation of ceramide. Direct in vitro measurement of enzymatic activity within CD95-stimulated U937 cell extracts, using labeled SM vesicles as substrates, showed strong SMase activity, which required pH 5.0 for optimal substrate hydrolysis. Finally, all CD95-sensitive cell lines tested could be induced to undergo apoptosis after exposure to cell-permeant C2-ceramide. These data indicate that CD95 cross-linking induces SM breakdown and ceramide production through an acidic SMase, thus providing the first information regarding early signal generation from CD95, and may be relevant in defining the biochemical nature of intracellular messengers leading to apoptotic cell death.
Article
Full-text available
Triggering of the Fas/APO-1 cell-surface receptor induces apoptosis through an uncharacterized chain of events. Exposure of Fas-sensitive cells to an agonist monoclonal antibody induced cell death and a 200-300% elevation in endogenous levels of the sphingolipid ceramide, a proposed intracellular mediator of apoptosis. In contrast, similar treatment of Fas-resistant cells caused insignificant changes in ceramide levels. Because resistant cell lines expressed the Fas antigen, these results indicate that these cells have a defect in the proximal signaling events leading to ceramide generation. Exposure of the resistant cell lines to a synthetic analog of ceramide induced apoptosis, thus bypassing Fas resistance and indicating that the signaling pathways downstream of ceramide were intact. Furthermore, activation of protein kinase C with the diacylglycerol analog phorbol 12-myristate 13-acetate significantly reduced Fas-induced cytotoxicity, suggesting opposing roles for ceramide and protein kinase C in regulation of apoptosis. These results provide evidence for ceramide as a necessary and sufficient lipid mediator of Fas-mediated apoptosis and suggest this process may be modulated via activation of additional signal-transduction pathways.
Article
Full-text available
Ceramide, a lipid mediator, has been most closely associated with antiproliferative activities. In this study, we examine the mechanism by which ceramide induces growth suppression and the role of the retinoblastoma gene product (Rb) in this process. Withdrawal of serum from the serum-dependent MOLT-4 cells resulted in significant dephosphorylation of Rb, correlating with the induction of G0/G1 cell cycle arrest. Serum withdrawal resulted in marked elevation in the levels of endogenous ceramide (3-fold at 24 h and 8-fold at 96 h) with little changes in the endogenous levels of sphingosine. The addition of exogenous C6-ceramide resulted in a concentration- and time-dependent dephosphorylation of Rb. Exogenous ceramide was active at levels comparable to endogenous levels achieved with serum withdrawal. Peak activity of exogenous ceramide (at 6 h) correlated with the uptake of C6-ceramide by MOLT-4 cells. Next, a number of studies were conducted to determine whether Rb plays a role in ceramide-induced growth suppression. (i) C6-Ceramide was poorly active in growth suppression of retinoblastoma cells that lack Rb. (ii) Mink lung epithelial cells in which Rb had been sequestered by overexpression of large tumor antigen were resistant to the action of ceramide compared to cells transfected with large tumor antigen mutated in the Rb-binding pocket. (iii) Overexpression of the EIA adenoviral protein, which binds and sequesters Rb, resulted in protection from growth suppression and cell cycle arrest induced by ceramide. Thus, these studies demonstrate that Rb is a downstream target for ceramide and may function in a growth suppressor pathway resulting in cell cycle arrest.
Article
Full-text available
Ceramide is emerging as a putative second messenger mediating effects of extracellular agents on cell growth and differentiation (Okazaki, T., Bielawska, A., Bell, R. M., and Hannun, Y. (1990) J. Biol. Chem. 265, 15822-15831) and programed cell death (Obeid, L. M., Linardic, C. M., Karolak, L. A., and Hannun, Y. A. (1993) Science 259, 1769-1771). In this study, the eight stereoisomers of C2-ceramide and dihydroceramide were synthesized, and their cellular activity was investigated. The four stereoisomers of C2-ceramide were active in inhibition of cell growth and induction of apoptosis with modest differences in potency. On the other hand, with C2-dihydroceramide only the threo compounds were active in these assays whereas the erythro compounds were totally inactive. Thus, of the two naturally occurring molecules, the analog of D-erythro-ceramide (with the 4-5 trans double bond) was active, whereas the analog of D-erythro-dihydroceramide was inactive. These results demonstrate the specificity of ceramide action and suggest that the introduction of the double bond is critical for imparting the biochemical and biological activity of ceramide.
Article
The retinoid, N-(4-hydroxyphenyl)retinamide (4-HPR), mediates p53-independent cytotoxicity and can increase reactive oxygen species and ceramide In solid tumor cell lines. We determined changes In ceramide and cytotoxicity upon treatment with 4-HPR (3-12 muM) In six human acute lymphoblastic leukemia (ALL) cell lines: T cell (MOLT-3, MOLT-4, CEM), pre-B-cell (NALM-6, SMS-SB), and null cell (NALL-1). Exposure to 4-HPR (12 muM) for 96 h caused 4.7 (MOLT-3), 3.5 (MOLT-4), 3.9 (CEM), 2.9 (NALM-6), 4.7 (SMS-SB), AND 4.5 (NALL-1) logs of cell kill. The average 4-HPR concentration that killed 99% of cells (LCgg) for all six lines was 4.8 muM (range: 1.5-8.9 muM). Treatment with 4-HPR (9 muM) for 24 h resulted in an 8.9 +/- 1.0-fold (range: 4.9-15.7-fold) increase of ceramide. Ceramide increase was time- and dose-dependent and abrogated by inhibitors of do novo ceramide synthesis. Concurrent inhibition of ceramide glycosylation/acylation by d,1-threo-(1-phenyl-2-hexadecanoyl-amino-3-morpholino-1-propanol) (PPMP) further increased ceramide levels, and synergistically increased 4-HPR cytotoxicity in four of six ALL cell lines. 4-HPR was minimally cytotoxic to peripheral blood mononuclear cells and a lymphoblastoid cell line, and increased ceramide <2-fold. Thus, 4-HPR was cytotoxic and increased ceramide in ALL cell lines, but not in nonmalignant lymphoid cell types.
Article
We previously reported that all-trans retinoic acid (RA) and fenretinide (4HPR) suppress HL-60 leukemia cell growth and cause partial cell arrest in the G1-to-S phase. Moreover, 4HPR but not RA induces apoptosis in HL-60 cells. To investigate further the observed biological effects, cyclin D1 and cdk4 expression and the level of phosphorylation of the retinoblastoma protein Rb were assessed. Cyclin D1 and cdk4 expression and Rb phosphorylation were significantly reduced, by 40-75%, after 24 hr of treatment with RA or 4HPR; these decreases were either transient, e.g., only at 24 hr for cdk4, or sustained for 72 hr. In general, more pronounced decreases were seen in the 4HPR-treated cells. Evidence for 4HPR-induced apoptosis comes from (1) cleavage of the enzyme poly(ADP-ribose) polymerase (PARP) to an 89-kDa truncated product, (2) appearance of DNA ladders on agarose gel electrophoresis, and (3) higher incorporation in situ of digoxigenin nucleotides into the free 3'-ends of DNA. Overnight pretreatment with 0.5-5.0 microM of the CPP32 inhibitor DEVD, but not the ICE inhibitor YVAD, significantly reduced the specific processing of PARP, suggesting that CPP32 is involved in the mechanism of action of 4HPR. Analysis of 2 lipid-derived second messengers, ceramide and diacylglycerol (DAG), as a function of time of treatment with RA or 4HPR, showed ceramide but not DAG to be significantly albeit transiently increased 2-fold at 3 hr, by 4HPR. To test further whether ceramide may be involved in the signaling cascade that culminates in the induction of apoptosis in 4HPR-treated HL-60 cells, the effects of fumonisin B1, an inhibitor of ceramide synthase, were studied. Simultaneous treatment of cells with 4HPR and 25-100 microM fumonisin B1 resulted in a dose-dependent reduction in the elevation in ceramide, the extent of PARP cleavage, and induction of apoptosis. Pretreatment with DEVD or YVAD, on the other hand, had no effect on the 4HPR-induced increase in ceramide.
Article
Several extracellular agents and stress stimuli, such as tumour necrosis factor alpha, chemotherapeutic agents and heat, cause ceramide accumulation. They do this by regulating enzymes involved in its metabolism. Ceramide modulates a number of biochemical and cellular responses to stress, including apoptosis, cell-cycle arrest and cell senescence.
Article
Lung cancer is the major cause of cancer-related death in the United States, with small-cell lung cancer (SCLC) constituting approximately 20% of all cases of lung cancer. Numerous epidemiologic and molecular studies have suggested that alterations in retinoid-signaling pathways play a role in the pathogenesis of lung cancer. Fenretinide [N-(4-hydroxyphenyl)retinamide; HPR] is a synthetic retinoid with minimal toxicity and favorable pharmacokinetics during long-term administration to patients in clinical trials. The aim of this investigation was to study the effect of HPR on the growth of SCLC cells in vitro. Seven SCLC cell lines (NCI-H69, NCI-H82, NCI-H146, NCI-H209, NCI-H345, NCI-H446, and NCI-H510A) were exposed continuously to a broad range of concentrations of HPR or all-trans-retinoic acid (RA), and cell viability was determined on day 3 and day 7 by the trypan blue dye exclusion assay. The growth of these cells was compared with that of control vehicle-treated cells to determine survival fraction and the dose resulting in a 50% inhibition of growth when compared with growth of control cells (IC50). The induction of apoptosis was evaluated by fluorescent microscopy, DNA content analysis, and a terminal deoxyribonucleotidyl transferase-based assay that labels 3'-hydroxyl ends of DNA fragments (TUNEL assay) combined with flow cytometric analysis. HPR inhibited growth of a panel of SCLC cell lines at IC50 values that ranged from 0.1 to 3.0 microM (concentrations that are clinically achievable). In all cell lines tested, HPR was a more potent growth inhibitor than RA. By use of fluorescent microscopy, HPR was found to induce morphologic changes consistent with apoptosis in NCI-H82 SCLC cells, including cellular shrinkage, chromatin condensation, and nuclear fragmentation. Flow cytometric analysis revealed decreased DNA content, and TUNEL assay showed increased digoxigenin-uridine triphosphate incorporation in HPR-treated NCI-H82 SCLC cells; these findings are consistent with the induction of apoptosis. HPR inhibited the in vitro growth of SCLC cells. In NCI-H82 cells, HPR inhibited growth via the induction of apoptosis.
Article
A proposed weak point in cancer cells is their need to synthesize novel or rare glucosphingolipids. It is further proposed that cancer patients be treated with a drug that slows the synthesis of glucosylceramide, the precursor of a large family of glucosphingolipids. Experimental data are furnished for chemotherapeutic and biochemical effects of PDMP, an analog of glucosylceramide and its precursor, ceramide. Promising results were obtained in the treatment of mice carrying Ehrlich ascites carcinoma cells and rats carrying C6 glioma cells. PDMP was found to be oxidized by cytochrome P-450, but this process could be blocked in vivo with piperonyl butoxide or cimetidine. A high level of blood glucose was found to elevate the size of rat kidneys and their content of UDP-glucose and its product, glucosylceramide. The excessive growth could be blocked by PDMP, which competes with UDP-glc for binding to glucosylceramide synthase. It is suggested that cancer patients be maintained at a low glucose level in order to slow the synthesis of glucosylceramide by tumor cells. Metabolic changes produced by PDMP in cultured cells, besides a rapid deletion of glucosphingolipids, were accumulation of the precursors (ceramide and sphingosine), loss of protein kinase C, and accumulation of diacylglycerol. It is suggested that many of the cellular changes produced by PDMP, such as loss of cell binding, are owing to existence of glucosylceramide-based "islands" floating in the outer cell surface; the islands may contain growth factor receptors and adhesion factors. An inhibitor that blocks sphingolipid synthesis, such as cycloserine, may prove to be a useful adjuvant for therapy with PDMP.