ArticlePDF Available

Kaempferol induced apoptosis via endoplasmic reticulum stress and mitochondria‐dependent pathway in human osteosarcoma U‐2 OS cells

Authors:

Abstract and Figures

Kaempferol is a natural flavonoid. Previous studies have reported that kaempferol has anti-proliferation activities and induces apoptosis in many cancer cell lines. However, there are no reports on human osteosarcoma. In this study, we investigate the anti-cancer effects and molecular mechanisms of kaempferol in human osteosarcoma cells. Our results demonstrate that kaempferol significantly reduces cell viabilities of U-2 OS, HOB and 143B cells, especially U-2 OS cells in a dose-dependent manner, but exerts low cytotoxicity on human fetal osteoblast progenitor hFOB cells. Comet assay, DAPI staining and DNA gel electrophoresis confirm the effects of DNA damage and apoptosis in U-2 OS cells. Flow cytometry detects the increase of cytoplasmic Ca(2+) levels and the decrease of mitochondria membrane potential. Western blotting and fluorogenic enzymatic assay show that kaempferol treatment influences the time-dependent expression of proteins involved in the endoplasmic reticulum stress pathway and mitochondrial signaling pathway. In addition, pretreating cells with caspase inhibitors, BAPTA or calpeptin before exposure to kaempferol increases cell viabilities. The anti-cancer effects of kaempferol in vivo are evaluated in BALB/c(nu/nu) mice inoculated with U-2 OS cells, and the results indicate inhibition of tumor growth. In conclusion, kaempferol inhibits human osteosarcoma cells in vivo and in vitro.
Content may be subject to copyright.
RESEARCH ARTICLE
Kaempferol induced apoptosis via endoplasmic
reticulum stress and mitochondria-dependent pathway
in human osteosarcoma U-2 OS cells
Wen-Wen Huang
1
, Yu-Jen Chiu
2
, Ming-Jen Fan
3
, Hsu-Feng Lu
4,5
, Hsiu-Feng Yeh
6
,
Kun-Hong Li
2
, Po-Yuan Chen
1
, Jing-Gung Chung
1,3
and Jai-Sing Yang
7
1
Department of Biological Science and Technology, China Medical University, Taichung, Taiwan
2
School of Medicine, China Medical University, Taichung, Taiwan
3
Department of Biotechnology, Asia University, Wufeng, Taichung, Taiwan
4
Department of Restaurant, Hotel and Institutional Management, Fu-Jen Catholic University, Taipei, Taiwan
5
Department of Clinical Pathology, Cheng-Hsin General Hospital, Taipei, Taiwan
6
School of Pharmacy, China Medical University, Taichung, Taiwan
7
Department of Pharmacology, School of Medicine, China Medical University, Taichung, Taiwan
Received: January 6, 2010
Revised: February 23, 2010
Accepted: March 25, 2010
Kaempferol is a natural flavonoid. Previous studies have reported that kaempferol has anti-
proliferation activities and induces apoptosis in many cancer cell lines. However, there are no
reports on human osteosarcoma. In this study, we investigate the anti-cancer effects and
molecular mechanisms of kaempferol in human osteosarcoma cells. Our results demonstrate
that kaempferol significantly reduces cell viabilities of U-2 OS, HOB and 143B cells, espe-
cially U-2 OS cells in a dose-dependent manner, but exerts low cytotoxicity on human fetal
osteoblast progenitor hFOB cells. Comet assay, DAPI staining and DNA gel electrophoresis
confirm the effects of DNA damage and apoptosis in U-2 OS cells. Flow cytometry detects the
increase of cytoplasmic Ca
21
levels and the decrease of mitochondria membrane potential.
Western blotting and fluorogenic enzymatic assay show that kaempferol treatment influences
the time-dependent expression of proteins involved in the endoplasmic reticulum stress
pathway and mitochondrial signaling pathway. In addition, pretreating cells with caspase
inhibitors, BAPTA or calpeptin before exposure to kaempferol increases cell viabilities. The
anti-cancer effects of kaempferol in vivo are evaluated in BALB/c
nu/nu
mice inoculated with
U-2 OS cells, and the results indicate inhibition of tumor growth. In conclusion, kaempferol
inhibits human osteosarcoma cells in vivo and in vitro.
Keywords:
Apoptosis / Endoplasmic reticulum stress / Kaempferol / Mitochondria-dependent/
U-2 OS
1 Introduction
Flavonoids are a class of plant secondary metabolites,
assorted into flavones, flavonols, flavanones, isoflavones,
and anthocyanidins [1]. It has been reported that intake of
flavonoids is associated with many biological properties,
such as antiviral [2], antitumor [3], anti-oxidative [4],
anti-inflammatory [5], hepatoprotective activities [6] and the
Correspondence: Professor Jing-Gung Chung, Department of
Biological Science and Technology, China Medical University,
No 91, Hsueh-Shih Road, Taichung City 404, Taiwan
E-mail: jgchung@mail.cmu.edu.tw
Fax: 1886-4-2205-3764
Abbreviations: PI, propidium iodide; DMSO, dimethyl sulfoxide;
FBS, fetal bovine serum; Z-LEHD-FMK, z-Leu-Glu-His-Asp-fluoro-
methyl ketone; Z-DEVDFMK, z-Asp-Met-Gln-Asp-fluoromethyl
ketone; hFOB, human fetal osteoblast; MTT, 3-(4,5-dimethyl-
thiazol-2-yl)-2,5-diphenyltetrazolium bromide; Dcm, mitochon-
drial membrane potential; ER, endoplasmic reticulum; EDTA,
ethylenediaminetetraacetic acid; DAPI, 4’,6-diamidino-2-phenyl-
indole
Additional corresponding author: Assistant Professor Jai-Sing Yang
E-mail: jaising@mail.cmu.edu.tw
&2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.mnf-journal.com
Mol. Nutr. Food Res. 2010, 54, 1585–1595 1585DOI 10.1002/mnfr.201000005
prevention of cardiovascular diseases [7]. Kaempferol, 3, 4’,
5, 7-tetra hydroxyflavone, a natural flavonoid, has been
isolated from various plant sources [8]. Kaempferol is
commonly known for antioxidant activity and is used for
cyto-protection agents. Previous studies have reported that
kaempferol has anti-proliferation activity and induces
apoptosis in various human cancer cell lines in vitro, such as
non-small cell lung cancer [9], leukemia [10], esophageal
cancer [11], prostate cancer [12], oral cavity cancer [13] and
colon cancer [14], but no reports on osteosarcoma.
High-grade osteosarcoma is the most common bone
malignancy, accounting for about 60% of malignant bone
tumors diagnosed in the first two decades of life, with an
aggressive local pattern of growth and high metastatic
potential [15]. Current standard treatment is to use chemo-
therapy followed by surgical resection [16]. The survival rate
of patients with localized osteosarcoma is about 11% with
surgery alone, compared to approximately 70% when
combined with chemotherapy [17]. Despite the success of
frontline therapy, about 40% of patients have progression
and further therapy with additional chemotherapy is
palliative and toxic. It is estimated that less than 30% of
patients with recurrent metastasis will be cured [18].
Chemotherapy-resistant cancer is one of the most serious
obstacles. Therefore, in this study, we focus on identifying
new agents to treat osteosarcoma.
Apoptosis is the process of programmed cell death that
occur in multi-cellular organisms, playing an important role
in normal physiology in animals [19]. However, impairment
of apoptotic function has been associated with several
diseases [20], such as neurodegenerative disorders and
cancers [21]. The perturbation of this process is considered a
crucial part of cancer prevention and therapy.
To date, one of the most effective anti-cancer strategies is
through the induction of apoptosis. Although previous
studies have reported that kaempferol has anti-cancer
activity in various human cancer cell lines [9–14], little is
known of the mechanisms exerting cytotoxicity in human
osteosarcoma. Therefore, the purpose of this study is to
investigate the apoptotic effects and the molecular
mechanisms of kaemperol in osteosarcoma cell lines in vitro
and in vivo.
2 Materials and methods
2.1 Chemicals and reagents
Kaempferol, propidium iodide (PI), Tris-HCl, calpeptin and
Triton X-100 were obtained from Sigma Chemical Co.
(St. Louis, MO, USA). BAPTA, dimethyl sulfoxide (DMSO)
and potassium phosphates were purchased from Merck Co.
(Darmstadt, Germany). Eagle’s minimum essential medium
(MEM), penicillin-streptomycin, trypsin-EDTA, fetal bovine
serum (FBS) and glutamine were obtained from Gibco BRL
(Grand Island, NY, USA). Caspase activity assay kit was
bought from OncoImmunin (MD, USA). Caspase-9 inhi-
bitor z-Leu-Glu-His-Asp-fluoromethyl ketone (Z-LEHD-
FMK) and caspase-3 inhibitor z-Asp-Met-Gln-Asp-fluoro-
methyl ketone (Z-DEVDFMK) were bought from R&D.
2.2 Human osteosarcoma cell lines (U-2 OS, HOB,
143B) and human fetal osteoblast progenitor cell
line (hFOB)
Human osteosarcoma U-2 OS, HOB, 143B cells and
conditionally immortalized human fetal osteoblast
progenitor hFOB cells were purchased from American Type
Culture Collection (ATCC). U-2 OS cells were cultured in
McCoy’s 5A medium (GIBCO-BRL) with 10% FBS and
antibiotics (100 U/ml of penicillin G and 100 mg/ml of
streptomycin) at 371C in a humidified atmosphere of 5%
CO
2
/95% air. HOB cells were cultured in minimum
essential medium (GIBCO-BRL) with 10% FBS and
antibiotics (100 U/ml of penicillin G and 100 mg/ml of
streptomycin) and 1.5 g/L sodium bicarbonate, 0.1 mM non-
essential amino acids at 371C in a humidified atmosphere of
5% CO
2
/95% air. 143B cells were cultured in minimum
essential medium (GIBCO-BRL) with 10% FBS and anti-
biotics (100 U/ml of penicillin G and 100 mg/ml of strepto-
mycin) and 0.015 mg/ml 5-bromo-20-deoxyuridine at 371Cin
a humidified atmosphere of 5% CO
2
/95% air. hFOB cells
were cultured in DMEM/Ham’s F12 medium (GIBCO-BRL)
with 10% FBS and antibiotics (100 U/ml of penicillin G,
100 mg/ml of streptomycin and 300 mg/ml geneticin) at
33.51C in a humidified atmosphere of 5% CO
2
/95% air [22].
2.3 Cell viability assay
The cell viability was determined by 3-(4,5-dimethylthiazol-
2-yl)-2,5- diphenyltetrazolium bromide (MTT) assay.
Human osteosarcoma U-2 OS, HOB, 143B cells and
conditionally immortalized human fetal osteoblast
progenitor hFOB cells were cultured in 96-well culture
plates and allowed to attach for hours before treated with
various concentrations (0, 25 50, 100, 150 or 200 mM) of
kaempferol. After cultivation for 24 h, 0.5 mg/ml of MTT
was then added to each well and the mixture was incubated
for 4 h at 371C. Culture medium was then replaced with an
equal volume of 0.04N HCl/isopropanol to dissolve forma-
zan crystals. Absorbance of each well was determined at
570 nm wavelength using ELISA reader [23].
2.4 Phase-contrast microscopy of morphological
changes
U-2 OS cells were plated in 24-well plates at a density of
2.5 10
5
cells/well. The 50, 100 or 150 mM of kaempferol
were added, and the cells were incubated for 24 h. A phase-
1586 W.-W. Huang et al.Mol. Nutr. Food Res. 2010, 54, 1585–1595
&2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.mnf-journal.com
contrast microscope was used for photography to determine
morphological changes as described elsewhere [24].
2.5 Comet assay and DAPI staining
After treated with 50, 100 or 150 mM of kaempferol, U-2 OS
cells were harvested and mixed with low melting point
(LMP) agarose at 371C. This mixture was placed on the top
of previous layer of 5% agarose (normal melting point) on
the slide, and then covered with a covership at 41C until
solid. Subsequently, the covership was removed gently and
some agarose was added onto this slide, and then covered
with the covership again. The slide was placed at 41C until
the mixture was solid, and put in chilled alkaline lysis buffer
for electrophoresis. Afterwards, the slide was gently washed
with neutralized buffer, and stained with DAPI [25].
2.6 Agarose gel electrophoresis
After treated with 150 mMofkaempferol,U-2OScellswere
harvested then lysed in lysis buffer (20 mM Tris, 10 mM
EDTA, 0.2 % Triton X-100, pH 8.0) at 41C for 15 min, and then
the lysate was centrifuged for 13,000 rpm, 10 min at 41C. The
supernatant containing fragmented DNA was collected and
incubated at 501C overnight with proteinase K (0.1 mg/ml) to
digest protein, followed by RNase A (50 mg/ml) digestion at
371C for 30 min. After extracted with phenol/chloroform/
isoamyl alcohol (25:24:1), the DNA was precipitated in 50%
isopropanol with 1 mlofglycogen(20mg/ml) at 201Cover-
night. The precipitated DNA was centrifuged at 14,000 rpm for
30 min, dried, and dissolved in 10 mlH
2
O. After electrophor-
esis in a 1.5% agarose gel containing ethidium bromide
(0.5 mg/ml) in TAE buffer (40 mM Tris-acetate, 1 mM EDTA,
pH 8.0), the DNA in gel was resolved with UV light [26].
2.7 Terminal deoxynucleotidyl transferase dUTP
nick end labeling (TUNEL) assay
After treated with 150 and 200 mM of kaempferol, U-2 OS
and hFOB cells were harvested. Evaluation of apoptosis in
the U-2 OS and hFOB cells were accomplished by flow
cytometry to detect cells labeled by TUNEL, using fluor-
escein-labeled dUTP (treatment). Controls consisted of cells
incubated with fluorescein dUTP without Tdt (In Situ Death
Kit, Boehringer-Mannheim Biochemicals) [27].
2.8 Intracellular Ca21levels assay
U-2 OS cells (2.5 10
5
/well) in 12-well plate were treated
with 150 mM of kaempferol and incubated for 0, 6, 12 or
24 h. Cells were harvested, washed twice, re-suspended in
3mg/ml of Indo 1/AM (Calbiochem; La Jolla, CA) at 371C for
30 min and analyzed by flow cytometry (Becton Dickinson
FACS Calibur) [26].
2.9 Calpain activity assays
U-2 OS cells were prepared on 24-well plates and pretreated
with BAPTA, a Ca
21
chelator or calpeptin and an inhibitor
of calpain for 1 h. Then, cells were loaded with 40 M Suc-
Leu-Leu-Val-Tyr-AMC calpain protease substrate (Biomol)
and treated with 150 mM of kaempferol to the indicated time
at 371C in a humidified 5% CO
2
incubator. Proteolysis of the
fluorescent probe was monitored by a fluorescent plate
reading system (HTS-7000 Plus Series BioAssay, Perkin
Elmer) with filter settings of 360720 nm for excitation and
460720 nm for emission [28].
2.10 Determination of mitochondrial membrane
potential (Dcm)
The mitochondrial membrane potential (Dc
m
) of the U-2
OS cells was determined by flow cytometry using DiOC6
(Molecular Probes). U-2 OS cells were treated with 150 mM
of kaempferol for 0, 6, 12 or 24 h to detect the changes of
Dc
m
. The cells were harvested and washed twice, re-
suspended in 500 ml of DiOC6 (4 mmol/L) and incubated at
371C for 30 min before analyzed by flow cytometry (Becton
Dickinson FACSCalibur) [29].
2.11 Caspase-3, -8 and -9 activities assay
Caspase-3, -8 and -9 activities were assessed according to
manufacturer’s instruction of caspase colorimetric kit (R&D
system Inc., MN, USA). U-2 OS cells were seeded in 12-well
cell culture plates at an initial density of 5.0 10
6
cells and
pretreated with caspase-3 inhibitor (Z-DEVD-FMK), caspase-
8 inhibitor (Z-IETD-FMK) or caspase-9 inhibitor (Z-LEHD-
FMK) for 1 h prior to treatment with 150 mM of kaempferol
for 0, 6, 12 or 24 h. Cells were harvested and lysed for 10 min
in 50 ml lysis buffer which contained 2 mM DTT.
After centrifugation, the supernatant containing 100 mg
protein were incubated with caspase-3 substrate (Ac-DEVD-
pNA), caspase-8 substrate (Ac-IETD-pNA) and caspase-9
substrate (Ac-LEHD-pNA) respectively in reaction buffer.
Then all samples were incubated in 96-well flat bottom
microplate at 371C for 1 h. Levels of released pNA
were measured at O.D.405 nm with ELISA reader (Anthos
2001) [23].
2.12 Western blot analysis
Briefly, the cytosolic and total proteins were collected from
U-2 OS cells which were treated with 150 mM of kaempferol
&2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.mnf-journal.com
Mol. Nutr. Food Res. 2010, 54, 1585–1595 1587
for 0, 6, 12 or 24 h. Protein concentrations were determined
by the Bio-Rad Protein Assay kit (Bio-Rad, Hercules, CA,
USA). Protein samples (30 mg each) were boiled with gel
loading buffer for 5 min. Protein extracts were separated on
10% SDS-polyacrylamide electrophoresis gels (SDS-PAGE)
and transferred onto a polyvinylidene difluoride (PVDF)
membrane. After blocking with TBST (0.05% Triton X-100
in PBS ) buffer of 5% non-fat milk for 1 h, the membrane
was exposed to the primary antibody: GADD153, GRP78,
GRP94, ATF-6a, ATF-6b, calpain 1, calpain 2, Fas, FasL,
Bax, Bcl-2, cytochrome c, Apaf-1, AIF, caspase-4, caspase-9,
caspase-3, caspase-7, caspase-8, caspase-12 and b-actin,
primary antibodies were diluted in PBST (0.05% Triton
X-100 in PBS ) buffer of 5% non-fat milk, and incubated at
41C overnight. The secondary antibodies were coupled
to horseradish peroxidase. Finally, they were detected by
ECL [27].
2.13 In vivo tumor xenograft model
Fifteen BALB/c
nu/nu
mice eight-week-old (approximately
22–28 g) were purchased from the Laboratory Animal Center,
National Taiwan University, College of Medicine (Taipei,
Taiwan). U-2 OS cells (1 10
7
) in culture medium were
subcutaneously injected into the flank of each mouse. Mice
with tumors were randomly assigned to three groups and
each group contained five animals. The treatment was initi-
ated when xenografts reached a volume of about 100 mm
3
and
these mice were treated orally every day with olive oil (control
vehicle), 25 mg/kg or 50 mg/kg of k aempferol in olive oil.
After xenograft tumor transplantation, mice were closely
monitored, counted and weighted. The tumor sizes were
measured every four days using calipers and tumor volume
was estimated according to the following formula: tumor
volume (mm
3
)5LxW
2
/2 (L: length and W: width). At the end
of the study, animals were sacrificed. Tumors were removed,
measured and weighted individually [29, 30].
2.14 Densitometry and statistical analysis
All data were expressed as mean7SEM from at least three
separate experiments. Statistical calculations of the data
were performed using an unpaired Student’s t-test. Statis-
tical significance was set at
Po0.05;

Po0.01;

Po0.001
was taken as significant.
3 Results
3.1 Effects of kaempferol on cell viability in human
osteosarcoma U-2 OS, HOS and 143B cells
We treated human osteosarcoma U-2 OS, HOB, 143B
and human fetal osteoblast progenitor hFOB cells
with kaempferol at different concentrations from 0 to
200 mM for 24 h. The number of viable cells was counted by
MTT method. As shown in Fig. 1A, the viability
was significantly decreased in the kaempferol-treated
human osteosarcoma cells groups, but not in hFOB
cells (IC
50
4200 mM). The IC
50
for U-2 OS cells was
148.36 mM. This therefore indicated that kaempferol
reduced the proportion of viable osteogenic cancer cells in
dose-dependent manner, but with low toxicity to hFOB
cells.
3.2 Effects of kaempferol on cell morphological
changes, DNA damage and apoptosis in human
osteosarcoma U-2 OS cells
To investigate the occurrence of morphological changes
and DNA damage in human osteosarcoma cells, we
predominantly focused on U-2 OS cells and treated
them with kaempferol at different concentrations from 0 to
150 mM for 24 h. In Fig. 1B, morphological examinations
of U-2 OS cells showed the difference between the kaemp-
ferol-treated groups and the control. In the kaempferol-
treated groups, cancer cells were detached from the
surface and contained some debris, whereas the control
group was well spread with a flattened morphology. In
Fig. 1C–D, the data showed that U-2 OS cells induced DNA
fragmentation and DNA damage was determined by
DAPI staining and comet assay. In Fig. 1E, in order to
reconfirm the induction of DNA damage, we isolated
DNA from the cells after treatment with 150 mM kaempferol
for 24 h, and then they were harvested for DNA fragmen-
tation determination in DNA gel electrophoresis. The
results showed that kaempferol induced apoptosis because
of the occurrence of DNA ladder. We investigated whether
or not kaempferol induces U-2 OS cell death through an
apoptotic mechanism. TUNEL assay was used for the
detection of DNA fragmentation in apoptosis. In Fig. 1F,
compared with control cells, U-2 OS cell were treated with
kaempferol showed significant cell apoptosis. However,
hFOB cell were showed non-significant cell apoptosis. We
suggested that kaempferol represented a promising candi-
date as an anti-osteosarcoma drug with low toxicity to
normal cells.
3.3 Effects of kaempferol on the cytoplasmic Ca21
and mitochondria membrane potential (Dcm)
levels in human osteosarcoma U-2 OS cells
In order to elucidate the possible signaling pathways of
kaempferol-induced apoptosis in U-2 OS cells,we examined
intracellular Ca
21
levels and mitochondria membrane
potential by flow cytometry analysis. As shown in Fig. 2A
and Fig. 3A, U-2 OS cells were treated with 150 mM
of kaempferol for 24 h and this significantly increased
1588 W.-W. Huang et al.Mol. Nutr. Food Res. 2010, 54, 1585–1595
&2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.mnf-journal.com
cytoplasmic Ca
21
levels and decreased Dcmin time-depen-
dent manner. These results suggested that kaempferol-
induced apoptotic response might be mediated by endo-
plasmic reticulum stress and mitochondrial-dependent
apoptotic pathways.
3.4 Effects of kaempferol on the levels of
endoplasmic reticulum stress related proteins in
human osteosarcoma U-2 OS cells
To be more detail in the molecular mechanisms of
endoplasmic reticulum stress pathway, we investigated
these related protein levels: GADD153, GRP78, GRP94,
ATF-6a, ATF-6b, caspase-4, caspase-12, calpain 1 and
calpain 2 by western blotting. As shown in Fig. 2B,
kaempferol increased these protein levels in time-depen-
dent manner, but caspase-12 has no statistical influence.
These results suggested that kaempferol-induced apop-
tosis was mediated via endoplasmic reticulum stress
pathway.
3.5 Effects of kaempferol with BAPTA or Calpeptin
on the levels of cytoplasmic Ca21and Calpain
activity and cell viability in human
osteosarcoma U-2 OS cells
In order to confirm that kaempferol-induced apoptosis
was mediated by ER stress pathway, we pretreated U-2 OS
cells with BAPTA, a Ca
21
chelator or calpeptin, an inhibitor
of calpain, after exposure to kaempferol. As shown in
Figure 1. Effects of kaempferol on cell
viability and apoptosis in osteo-
sarcoma cell lines. After treatment with
various concentrations of kaempferol
for 24 h, the cell viabilities of U-2 OS,
HOB, 143B osteosarcoma cell lines and
the conditionally immortalized human
fetal osteoblast progenitor hFOB
cells are shown (A). Data represent
mean7SD of three experiments.

po0.001. U-2 OS cells in response
to various concentrations of kaemp-
ferol for 24 h showed morphological
changes (B) which indicated kaemp-
ferol-induced cell death, and DAPI
staining (C), comet assay (D), gel elec-
trophoresis (E) and TUNEL assay (F)
revealing kaempferol-induced DNA
damage, fragmentation and apoptosis,
which was another hallmark of cells
undergoing apoptosis.
Mol. Nutr. Food Res. 2010, 54, 1585–1595 1589
&2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.mnf-journal.com
Fig. 2C–D, the levels of calpain activity and cell viability
were significantly influenced. Overall, in Fig. 2, these data
demonstrated that activation of ER stress pathway played an
important role in kaempferol-induced apoptosis in U-2 OS
cells.
3.6 Effects of kaempferol on the levels of Bcl-2
family in human osteosarcoma U-2 OS cells
Previous studies have demonstrated that Bcl-2 and Bax
locate in the mitochondrial outer-membrane and the
Bcl-2/Bax ratio regulate the release of mitochondrial
cytochrome cto cytosol [31, 32]. We investigated expression
levels of Bcl-2 and Bax in kaempferol-treated U-2 OS cells by
western blotting. As shown in Fig. 3B, the pro-apoptotic
protein level of Bax was up-regulated, whereas the anti-
apoptotic protein level of Bcl-2 was down-regulated in time-
dependent manner.
3.7 Effects of kaempferol on the levels of
mitochondrial caspase-dependent and caspase-
independent pathway related proteins in human
osteosarcoma U-2 OS cells
To be more detail in the molecular mechanisms of mito-
chondrial-dependent apoptotic pathway, we examined the
expression levels of cytochrome c, Apaf-1, caspase-9, caspase-3,
caspase-7 and AIF by western blotting. As shown in Fig. 3B,
these protein levels were increased in time-dependent manner.
Our results suggested that kaempferol-induced apoptotic
response was mediated by mitochondrial-dependent cascade.
3.8 Effects of kaempferol on the caspase-9 and caspase
-3 activities in human osteosarcoma U-2 OS cells
In order to confirm that kaempferol-induced apoptosis was
mediated by caspase-dependent pathway, we investigated the
Figure 2. Effects of kaempferol
on U-2 OS cells in endoplasmic
reticulum stress apoptotic
pathway. The intracellular Ca
21
levels in kaempferol-treated
U-2 OS cells from each time
point were measured by flow
cytometric analysis (A). Cells
were treated with 150 mMof
kaempferol for the indicated
time, cytosolic proteins or
whole cell lysate were
prepared, and subjected to
Western blotting. The resulting
blots were probed for
GADD153, GRP78, GRP94, ATF-
6a,ATF-6b, caspase-4, caspase-
12, calpain1 and calpain2
(whole cell lysate). b-actin
served as the loading control.
Levels of the associated
proteins in endoplasmic reticu-
lum stress apoptotic pathway
were affected (B). Cells were
pretreated with BAPTA, a Ca
21
chelator or calpeptin, an inhi-
bitor of calpain for 1 h after
exposure to kaempferol, then
incubated for 24 h. The whole-
cell lysates were subjected to
calpain activity assay (C) and
cells were collected to deter-
mine the percentage of viable
cells (D). Data from three inde-
pendent experiments were
presented (

Po0.001, as
compared with control treat-
ments).
1590 W.-W. Huang et al.Mol. Nutr. Food Res. 2010, 54, 1585–1595
&2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.mnf-journal.com
caspase-9, -3 and -8 activities by fluorogenic enzymatic assay.
As shown in Fig. 3C–D, both caspase-9 and caspase-3 activ-
ities were significantly increased. Moreover, pre-incubation
with specific inhibitors of caspases-9 (Z-LEHD-FMK), caspase-
3 (Z-DEVE-FMK) strongly reduced the caspase-9 or caspase-3
activities and increased U-2 OS cell viability. However,
caspase-8 activity has no significant influence. Overall, in
Fig. 3, these data demonstrated that caspase-dependent
mitochondrial pathway played an important role in kaemp-
ferol-induced apoptosis in U-2 OS cells.
3.9 Effects of kaempferol on anti-proliferative
activity in BALB/c
nu/nu
mice after injection with
human osteosarcoma U-2 OS cells
Three groups of mice were respectively treated with DMSO
control vehicle, 25 mg/kg or 50 mg/kg of kaempferol. These
representative animals with tumors were shown in Fig. 4A.
In Fig. 4B–C, kaempferol significantly decreased the
tumor weight and tumor volume compared to the control
group.
Figure 3. Effects of kaempferol
on U-2 OS cells in mitochon-
drial-dependent apoptotic pa-
thway. The mitochondrial
membrane potential (Dc
m
)of
kaempferol-treated U-2 OS
cells from each time point was
measured by staining with
DiOC6 (A). Cells were treated
with 150 mM of kaempferol for
the indicated time, cytosolic
proteins or whole cell lysate
were prepared, and subjected
to Western blotting. The
resulting blots were probed for
cytochrome c, Apaf-1, AIF,
caspase-9, caspase-3, caspase-
7 (cytosolic proteins), and Bcl-
2,Bax (whole cell lysate).
b-actin served as the loading
control. Levels of the asso-
ciated proteins in mitochon-
drial-dependent apoptotic
pathway were affected (B).
Cells were pretreated with
specific inhibitors of caspases-
9 (Z-LEHD-FMK), caspase-3
(Z-DEVE-FMK) or caspase-8
inhibitor (Z-IETD-FMK) for 1 h
after exposure to kaempferol,
then incubated for 24 h.
The whole-cell lysates were
subjected to caspase activity
assay (C) and cells were
collected to determine the
percentage of viable cells (D).
Data from three independent
experiments were presented
(

Po0.001, as compared
with control treatments).
Mol. Nutr. Food Res. 2010, 54, 1585–1595 1591
&2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.mnf-journal.com
4 Discussion
Kaempferol, a natural flavonoid, has been reported to induce
apoptosis and inhibit proliferation in various human cancer
cell lines, including non-small cell lung cancer [9], leukemia
[10], esophageal cancer [11], prostate cancer [12], oral cavity
cancer [33] and colon cancer [14]. Furthermore, Zhang et al.
demonstrated that kaempferol not only effectively inhibited
pancreatic cancer cell proliferation and induced apoptosis,
but also may sensitized pancreatic tumor cells to chemo-
therapy [34]. However, little is known in human osteo-
sarcoma cell lines. In contrast to beneficial effects, there are
still some question marks about the toxic side-effects to
normal tissue. Li et al. showed cytotoxicities of kaempferol at
higher doses in human normal liver L-02 cells (IC
50
5
57.05 mM, cultivation for 48 h) and human hepatoma
HepG2 cells (IC
50
584.72 mM, cultivation for 48 h) in vitro
[35]. Soares et al. showed that the viability of kaempferol-
treated mouse fibroblast McCoy cells was fell, without the
hepatic S9 microsomal fraction; but low toxicity occurred
(IC
50
4500 mM) when the S9 mixture metabolized these
compounds [36]. In this study, we first reported that
kaempferol was active against human osteosarcoma U-2 OS,
HOB and 143B cell lines in vitro and U-2 OS in vivo.In
Fig. 1, it was shown that kaempferol reduced the percentage
of viable cancer cells in a dose-dependent manner and
induced apoptotic cell death in U-2 OS cells; however, it
exhibited low toxicity to human fetal osteoblast progenitor
hFOB cells (IC
50
4200 mM).
Inducing apoptosis in cancer cells is one of the major
strategies of cancer therapeutics. Three major pathways lead to
apoptosis [37]. First, the death receptor pathway is triggered by
the binding of extrinsic signals to surface receptors, resulting
in activation of caspase-8 followed by the activation of caspase-3
and -7 [19]. Second, the mitochondrial pathway is triggered by
various stimuli damage inside the cell. When an excess of pro-
apoptotic over anti-apoptotic signals, it initiates mitochondrial
outer membrane permeabilization and results in caspase
dependent and independent apoptotic pathway [31, 32]. Kang
et al. demonstrated that kaempferol and quercetin, compo-
nents of ginkgo biloba extract, induced caspase-3-dependent
apoptosis in oral cavity cancer cell lines, SCC-1483, SCC-25
and SCC-QLL1 [33]. Leung et al. showed that kaempferol-
induced apoptosis in human lung non-small carcinoma H460
cells was through caspase-3 (caspase-dependent) and AIF
(caspase-independent) pathways [9]. Zhang et al. reported that
kaempferol exerted cytotoxic effects on OE33, a human
esophageal adenocarcinoma cell line, causing G2/M arrest and
inducing caspase-dependent apoptosis [11]. Furthermore,
Marfe et al. demonstrated that kaempferol induced apoptosis
in K562 and U937 leukemia cell lines via Akt inactivation and
mitochondrial dysfunction [10]. Our results are in agreement
with previous studies. In Fig. 3, these data indicated that
kaempferol up-regulated the level of pro-apoptotic protein Bax
and down-regulated anti-apoptotic protein Bcl-2, accompanied
with the loss of Dc
m
, and then promoting activities of caspase-
9,-3,and-7,butnotcaspase-8.Specicinhibitorsofcaspase-9
and -3 which decreased caspase activities and increased the
kaempferol-treated cell viability suggested that kaempferol
induced apoptosis through the mitochondrial-dependent
pathway in U-2 OS cells. Also, up-regulating the protein level
of AIF indicated that apoptosis was also undergone via case-
pase-independent mitochondrial pathway. Third, the novel
endoplasmic reticulum (ER)-specific apoptotic pathway, it is
induced by accumulation of unfolded/misfolded protein
aggregating in ER or by excessive protein traffic. Increasing the
proteins level of GADD153, GRP78, GRP94 and ATF which
are the hallmarks of ER stress induces a rise in intracellular
Ca
21
level, mitochondrial membrane depolarization and
Figure 4. In vivo anti-tumor activity of kaempferol. BALB/c
nu/nu
mice were administered 25 and 50 mg/kg of kaempferol orally.
Representative animals with tumors (A), tumor weight (B) and
total tumor volume of BALB/c
nu/nu
mice (C). Data were presented
(

Po0.001, as compared with control treatments).
1592 W.-W. Huang et al.Mol. Nutr. Food Res. 2010, 54, 1585–1595
&2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.mnf-journal.com
activation of calpain and caspase-12 in murine systems and/or
caspase-4 in human cells [38–40]. However, there are no
reports about kaempferol-induced ER stress in cancer cells. In
Fig. 2, increased levels of GADD153, GRP78, GRP94, ATF-6a
and ATF-6bwere followed by releasing Ca
21
from ER,
increasing calpain proteins expression and activating caspase-
4, and finally leading to apoptosis. Our result was shown in
fig. 2B and caspase-12 protein expression level has no signif-
icance influence. Caspase-12 has been shown to be involved
ER stress-induced apoptosis pathways, but in humans,
although the caspase-12 gene is transcribed into mRNA,
mature caspase-12 protein would not be produced because the
gene is interrupted by a frame shift and a premature stop
codon. [38–40]. ER stress signaling pathway was reconfirmed
by pre-treating with BAPTA, a Ca
21
chelator, and calpeptin, an
inhibitor of calpain, in kaempferol-treated U-2 OS cells, and it
showed decrease of calpain activity and increase of cell viability.
These accumulating data demonstrated that the activation of
ER stress pathway played an important role in kaempferol-
induced apoptosis in U-2 OS cells.
Concentrated and selected accumulation of anti-cancer
drugs at the tumor site is essential for the success of drug
treatment in vivo. Previous studies have reported that flavonoid
exhibits ability to inhibit human colorectal tumor formation
and block rat glioma tumoral invasion and migration in vivo
[41, 42]. Besides, two cohort studies have showed that high
level of kaempferol intake significantly decreases ovarian
cancer incidence, and intake of flavonol and catechin may be
associated with a decreased colorectal cancer risk in normal
weight women [43, 44]. In Fig. 4, our results showed that both
25 mg/kg and 50 mg/kg of kaempferol significantly reduced
the tumor volume and weight in BALB/c
nu/nu
osteosarcoma
mice. Additional prospective studies are needed to further
evaluate these associations.
In conclusion, with this report, we now show that
kaempferol exhibits direct anti-tumor activity, inducing tumor
cell apoptosis and suppressing tumor cell proliferation.
Moreover, kaempferol induced apoptosis through the mito-
chondria- dependent and ER stress pathways in human
osteosarcoma U-2 OS cells. Finally, we show that kaempferol
profoundly suppresses the in U-2 OS tumor xenograft-bearing
mice in vivo. The proposed signal pathways of kaempferol-
induced apoptosis in human osteosarcoma U-2 OS cells are
shown in Fig. 5. Although there are still some controversy
about the safety and biological effects of flavonoids, these
findings provide important possible molecular mechanisms of
the anti-human osteosarcoma and confirm that kaempferol
may be an anti-osteosarcoma cancer drug candidate.
The investigation was supported by a research grant from
the National Science Council of the Republic of China (NSC
97-2320-B-039 -004 -MY3) and a grant from the China Medical
University (CMU94-056), Taiwan.
The authors have declared no conflict of interest.
5 References
[1] Hollman, P. C., Katan, M. B., Dietary flavonoids: intake,
health effects and bioavailability. Food Chem. Toxicol. 1999,
37, 937–942.
Figure 5. A proposed model illustrates the
molecular mechanism and the overall
possible signaling pathways of kaempferol-
induced apoptosis in U-2 OS cells.
Mol. Nutr. Food Res. 2010, 54, 1585–1595 1593
&2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.mnf-journal.com
[2] Guo, Q., Zhao, L., You, Q., Yang, Y. et al., Anti-hepatitis B
virus activity of wogonin in vitro and in vivo.Antiviral Res.
2007, 74, 16–24.
[3] Cardenas, M., Marder, M., Blank, V. C., Roguin, L. P., Anti-
tumor activity of some natural flavonoids and synthetic
derivatives on various human and murine cancer cell lines.
Bioorg. Med. Chem. 2006, 14, 2966–2971.
[4] Burda, S., Oleszek, W., Antioxidant and antiradical activities
of flavonoids. J. Agric. Food. Chem. 2001, 49, 2774–2779.
[5] Gonzalez-Gallego, J., Sanchez-Campos, S., Tunon, M. J.,
Anti-inflammatory properties of dietary flavonoids. Nutr.
Hosp. 2007, 22, 287–293.
[6] Yao, P., Nussler, A., Liu, L., Hao, L. et al., Quercetin protects
human hepatocytes from ethanol-derived oxidative stress
by inducing heme oxygenase-1 via the MAPK/Nrf2 path-
ways. J. Hepatol. 2007, 47, 253–261.
[7] Tijburg, L. B., Mattern, T., Folts, J. D., Weisgerber, U. M.,
Katan, M. B., Tea flavonoids and cardiovascular disease: a
review. Crit. Rev. Food. Sci. Nutr. 1997, 37, 771–785.
[8]Park,J.S.,Rho,H.S.,Kim,D.H.,Chang,I.S.,Enzymatic
preparation of kaempferol from green tea seed and its anti-
oxidant activity. J. Agric. Food. Chem. 2006, 54, 2951–2956.
[9] Leung, H. W., Lin, C. J., Hour, M. J., Yang, W. H. et al.,
Kaempferol induces apoptosis in human lung non-small
carcinoma cells accompanied by an induction of antioxidant
enzymes. Food Chem. Toxicol. 2007, 45, 2005–2013.
[10] Marfe, G., Tafani, M., Indelicato, M., Sinibaldi-Salimei, P.
et al., Kaempferol induces apoptosis in two different cell lines
via Akt inactivation, Bax and SIRT3 activation, and mito-
chondrial dysfunction. J. Cell. Biochem. 2009, 106, 643–650.
[11] Zhang, Q., Zhao, X. H., Wang, Z. J., Flavones and flavonols
exert cytotoxic effects on a human oesophageal adenocarci-
noma cell line (OE33) by causing G2/M arrest and inducing
apoptosis. Food Chem. Toxicol. 2008, 46, 2042–2053.
[12] De Leo, M., Braca, A., Sanogo, R., Cardile, V. et al., Anti-
proliferative activity of Pteleopsis suberosa leaf extract and
its flavonoid components in human prostate carcinoma
cells. Planta. Med. 2006, 72, 604–610.
[13] Kim, K. S., Rhee, K. H., Yoon, J. H., Lee, J. G. et al.,Ginkgo
biloba extract (EGb 761) induces apoptosis by the activation
of caspase-3 in oral cavity cancer cells. Oral Oncol. 2005, 41,
383–389.
[14] Mutoh, M., Takahashi, M., Fukuda, K., Matsushima-Hibiya, Y.
et al., Suppression of cyclooxygenase-2 promoter-dependent
transcriptional activity in colon cancer cells by chemopre-
ventive agents with a resorcin-type structure. Carcinogenesis
2000, 21,959963.
[15] Arndt, C. A., Crist, W. M., Common musculoskeletal tumors
of childhood and adolescence. N. Engl. J. Med. 1999, 341,
342–352.
[16] Bielack, S., Carrle, D., Jost, L., Osteosarcoma: ESMO clinical
recommendations for diagnosis, treatment and follow-up.
Ann. Oncol. 2008, 19, ii94–ii96.
[17] Siegel, H. J., Pressey, J. G., Current concepts on the surgical
and medical management of osteosarcoma. Expert Rev.
Anticancer Ther. 2008, 8, 1257–1269.
[18] Duan, Z., Choy, E., Harmon, D., Yang, C. et al., Insulin-like
growth factor-I receptor tyrosine kinase inhibitor cyclo-
lignan picropodophyllin inhibits proliferation and induces
apoptosis in multidrug resistant osteosarcoma cell lines.
Mol. Cancer. Ther. 2009, 8, 2122–2130.
[19] Degterev, A., Boyce, M., Yuan, J., A decade of caspases.
Oncogene 2003, 22, 8543–8567.
[20] Lockshin, R. A., Zakeri, Z., Cell death in health and disease.
J. Cell Mol. Med. 2007, 11, 1214–1224.
[21] Moran, J. M., Gonzalez-Polo, R. A., Ortiz-Ortiz, M. A., Niso-
Santano, M. et al., Identification of genes associated with
paraquat-induced toxicity in SH-SY5Y cells by PCR array
focused on apoptotic pathways. J. Toxicol. Environ. Health
A2008, 71, 1457–1467.
[22] Hsu, S. C., Yang, J. S., Kuo, C. L., Lo, C. et al., Novel
quinolone CHM-1 induces apoptosis and inhibits metastasis
in a human osterogenic sarcoma cell line. J. Orthop. Res.
2009, 27, 1637–1644.
[23] Lin, S. S., Huang, H. P., Yang, J. S., Wu, J. Y. et al., DNA
damage and endoplasmic reticulum stress mediated
curcumin-induced cell cycle arrest and apoptosis in human
lung carcinoma A-549 cells through the activation caspases
cascade- and mitochondrial-dependent pathway. Cancer
Lett. 2008, 272, 77–90.
[24] Ho, Y. T., Yang, J. S., Li, T. C., Lin, J. J. et al., Berberine
suppresses in vitro migration and invasion of human SCC-4
tongue squamous cancer cells through the inhibitions of
FAK, IKK, NF-kappaB, u-PA and MMP-2 and -9. Cancer Lett.
2009, 279, 155–162.
[25] Chou, L. C., Yang, J. S., Huang, L. J., Wu, H. C. et al., The
synthesized 2-(2-fluorophenyl)-6,7-methylenedioxyquino-
lin-4-one (CHM-1) promoted G2/M arrest through inhibition
of CDK1 and induced apoptosis through the mitochondrial-
dependent pathway in CT-26 murine colorectal adeno-
carcinoma cells. J. Gastroenterol 2009, 44, 1055–1063.
[26] Yang, J. S., Chen, G. W., Hsia, T. C., Ho, H. C. et al., Diallyl
disulfide induces apoptosis in human colon cancer cell line
(COLO 205) through the induction of reactive oxygen
species, endoplasmic reticulum stress, caspases casade
and mitochondrial-dependent pathways. Food Chem.
Toxicol. 2009, 47, 171–179.
[27] Chung, J. G., Yang, J. S., Huang, L. J., Lee, F. Y. et al.,
Proteomic approach to studying the cytotoxicity of YC-1
on U937 leukemia cells and antileukemia activity in ortho-
topic model of leukemia mice. Proteomics 2007, 7,
3305–3317.
[28] Lee, M. J., Kee, K. H., Suh, C. H., Lim, S. C., Oh, S. H.,
Capsaicin-induced apoptosis is regulated by endoplasmic
reticulum stress- and calpain-mediated mitochondrial cell
death pathways. Toxicology 2009, 264, 205–214.
[29] Wen, Y. F., Yang, J. S., Kuo, S. C., Hwang, C. S. et al.,
Investigation of anti-leukemia molecular mechanism of ITR-
284, a carboxamide analog, in leukemia cells and its effects
in WEHI-3 leukemia mice. Biochem. Pharmacol. 2010, 79,
389–398.
[30] Koch, T. C., Briviba, K., Watzl, B., Fahndrich, C. et al.,
Prevention of colon carcinogenesis by apple juice in vivo:
1594 W.-W. Huang et al.Mol. Nutr. Food Res. 2010, 54, 1585–1595
&2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.mnf-journal.com
impact of juice constituents and obesity. Mol. Nutr. Food
Res. 2009, 53, 1289–1302.
[31] Danial, N. N., Korsmeyer, S. J., Cell death: critical control
points. Cell 2004, 116, 205–219.
[32] Chen, J. Q., Brown, T. R., Yager, J. D., Mechanisms of
hormone carcinogenesis: evolution of views, role of mito-
chondria. Adv. Exp. Med. Biol. 2008, 630, 1–18.
[33] Kang, J. W., Kim, J. H., Song, K., Kim, S. H. et al., Kaemp-
ferol and quercetin, components of Ginkgo biloba extract
(EGb 761), induce caspase-3-dependent apoptosis in oral
cavity cancer cells. Phytother. Res. 2010, 148, S77–82.
[34] Zhang, Y., Chen, A. Y., Li, M., Chen, C., Yao, Q., Ginkgo
biloba extract kaempferol inhibits cell proliferation and
induces apoptosis in pancreatic cancer cells. J. Surg. Res.
2008, 148, 17–23.
[35] Li, N., Liu, J. H., Zhang, J., Yu, B. Y., Comparative evaluation
of cytotoxicity and antioxidative activity of 20 flavonoids.
J. Agric. Food Chem. 2008, 56, 3876–3883.
[36] Soares, V. C., Varanda, E. A., Raddi, M. S., In vitro basal and
metabolism-mediated cytotoxicity of flavonoids. Food
Chem. Toxicol. 2006, 44, 835–838.
[37] Ziegler, D. S., Kung, A. L., Therapeutic targeting of apoptosis
pathways in cancer. Curr. Opin. Oncol. 2008, 20, 97–103.
[38] Nieto-Miguel, T., Fonteriz, R. I., Vay, L., Gajate, C. et al.,
Endoplasmic reticulum stress in the proapoptotic action of
edelfosine in solid tumor cells. Cancer Res. 2007, 67,
10368–10378.
[39] Hitomi, J., Katayama, T., Eguchi, Y., Kudo, T. et al., Invol-
vement of caspase-4 in endoplasmic reticulum stress-
induced apoptosis and Abeta-induced cell death. J. Cell
Biol. 2004, 165, 347–356.
[40] Rao, R. V., Ellerby, H. M., Bredesen, D. E., Coupling endo-
plasmic reticulum stress to the cell death program. Cell
Death Differ. 2004, 11, 372–380.
[41] Chen, Y. C., Shen, S. C., Chow, J. M., Ko, C. H., Tseng, S. W.,
Flavone inhibition of tumor growth via apoptosis in vitro
and in vivo.Int. J. Oncol. 2004, 25, 661–670.
[42] Shen, S. C., Lin, C. W., Lee, H. M., Chien, L. L., Chen, Y. C.,
Lipopolysaccharide plus 12-o-tetradecanoylphorbol 13-
acetate induction of migration and invasion of glioma cells
in vitro and in vivo: Differential inhibitory effects of flavo-
noids. Neuroscience 2006, 140, 477–489.
[43] Gates, M. A., Tworoger, S. S., Hecht, J. L., De Vivo, I. et al.,
A prospective study of dietary flavonoid intake and inci-
dence of epithelial ovarian cancer. Int. J. Cancer 2007, 121,
2225–2232.
[44] Simons, C. C., Hughes, L. A., Arts, I. C., Goldbohm, R. A.
et al., Dietary flavonol, flavone and catechin intake and risk
of colorectal cancer in the Netherlands Cohort Study. Int.
J. Cancer 2009, 125, 2945–2952.
&2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.mnf-journal.com
Mol. Nutr. Food Res. 2010, 54, 1585–1595 1595
... These studies reported the apoptosis-inducing properties of KMP, which can be partially attributed to its impacts on the MAPK pathway. In human lung cancer cells (A549 cell line), activation of the mitogen-activated protein kinase (MAPK) pathway is a key factor in Most of the studies included in the present review explored the effects of KMP on tumor cell death [10,12,22,[27][28][29][30][31][32]35,36,40,46,[51][52][53][55][56][57]62,63,66,70,[74][75][76][77]80,81]. These studies reported the apoptosis-inducing properties of KMP, which can be partially attributed to its impacts on the MAPK pathway. ...
... The authors reported positive results, including reduced colony formation, cell proliferation, migration, and invasion and increased cell viability, apoptosis, and cell cycle arrest in the G1 phase. Huang et al. [35] studied the anticancer effects and molecular mechanisms of KMP in human osteosarcoma cells and demonstrated that KMP significantly reduced the viability of U-2 OS, HOB, and 143B cells in a dose-dependent manner, with low cytotoxicity on hFOB cells, a human fetal osteoblast progenitor cell line. In vitro assays confirmed the effects of DNA damage, apoptosis in U-2 OS cells, increased cytoplasmic levels of calcium ions, and decreased mitochondrial membrane potential [35]. ...
... Huang et al. [35] studied the anticancer effects and molecular mechanisms of KMP in human osteosarcoma cells and demonstrated that KMP significantly reduced the viability of U-2 OS, HOB, and 143B cells in a dose-dependent manner, with low cytotoxicity on hFOB cells, a human fetal osteoblast progenitor cell line. In vitro assays confirmed the effects of DNA damage, apoptosis in U-2 OS cells, increased cytoplasmic levels of calcium ions, and decreased mitochondrial membrane potential [35]. In additional experiments, Chen et al. [42] later showed that KMP decreased the DNA-binding activity of activator protein-1 transcription factor (AP-1), suggesting a potential role of the compound in the treatment of osteosarcoma metastasis. ...
Article
Full-text available
Simple Summary Kaempferol, a natural compound commonly found in fruits, vegetables, and plants, has gained interest within the scientific community because of its anticancer properties against different types of tumors. The results of this review reveal that kaempferol exerts anticancer effects on many types of tumor cells by different mechanisms, providing evidence of its potential as a cancer drug. Abstract Given the heterogeneity of different malignant processes, planning cancer treatment is challenging. According to recent studies, natural products are likely to be effective in cancer prevention and treatment. Among bioactive flavonoids found in fruits and vegetables, kaempferol (KMP) is known for its anti-inflammatory, antioxidant, and anticancer properties. This systematic review aims to highlight the potential therapeutic effects of KMP on different types of solid malignant tumors. This review was conducted following the Preferred Reporting Items for Systematic Review and Meta-Analyses (PRISMA) guidelines. Searches were performed in EMBASE, Medline/PubMed, Cochrane Collaboration Library, Science Direct, Scopus, and Google Scholar. After the application of study criteria, 64 studies were included. In vitro experiments demonstrated that KMP exerts antitumor effects by controlling tumor cell cycle progression, proliferation, apoptosis, migration, and invasion, as well as by inhibiting angiogenesis. KMP was also able to inhibit important markers that regulate epithelial–mesenchymal transition and enhanced the sensitivity of cancer cells to traditional drugs used in chemotherapy, including cisplatin and 5-fluorouracil. This flavonoid is a promising therapeutic compound and its combination with current anticancer agents, including targeted drugs, may potentially produce more effective and predictable results.
... Icariside II against osteosarcoma [26], cervical [27], breast [28], glioma [29], [30], esophageal [31], and glioma [32]. Gallic acid against Lung Cancer [33], Osteosarcoma [34], Myricetin against Bladder Cancer [35], prostate cancer [36] Caffeic Acid against Lung Cancer [37], Curcumin against glioma [38] and glioblastoma [39], Quercetin against cervical cancer [40], Sinapic acid against pancreatic cancer [41], Kaempferol against Gastric cancer [42], cholangiocarcinoma [43], Osteosarcoma [44], Pterostilbene against Breast cancer [45], non-small-cell lung cancer [46], Esophageal Cancer [47], Colon Cancer [48], Resveratrol against bladder cancer [49,50], colorectal cancer [51], lung cancer [52,53], Gnetin C against prostate cancer [54], epigallocatechin gallate ...
Chapter
Full-text available
Polyphenols, a wide group of secondary metabolites found in plants and some animals too, have gained significant attention in recent years due to their outstanding bioactivity and potential therapeutic applications. This chapter will provide a brief overview of the bioactivity of polyphenols, highlighting their importance in the treatment of different disorders. Polyphenols exhibit strong anticancer properties, demonstrated through multiple in vitro and in vivo studies. Their potential to inhibit cancer growth, progression, metastasis, induction of apoptosis, and modulation of multiple signaling pathways involved in cancer make them effective candidates for the treatment of cancer. Furthermore, polyphenols have significant antibacterial action, by disrupting cell membranes and inhibiting the synthesis of different enzymes, making them valuable agents for overcoming bacterial drug resistance. Moreover, they also possess antiaging properties, attributed to their strong antioxidant potential. They assist in combating cellular damage and reducing the aging process by lowering oxidative stress and scavenging free radicals. In addition to all this, polyphenols exert antidiabetic effects by modulating the metabolism of glucose, increasing the sensitivity of insulin, and lowering oxidative stress, offering potential therapeutic benefits for being effective against diabetes mellitus. Polyphenols also show cardioprotective effects, with evidence describing their potential to improve cardiovascular health by lowering inflammation, blood pressure, and free radicals, and inhibiting aggregation of platelets. Furthermore, recent studies also highlight antiviral, anti-Alzheimer, antifungal, and antiparasitic activities. In conclusion, the abundance of literature overwhelmingly demonstrates the bioactivity of polyphenols against various diseases. Understanding the primary mechanisms behind the bioactivity of polyphenols holds great promise for developing innovative therapeutic interventions for different disease conditions.
... Kaempferol diketahui dapat menginduksi apoptosis pada garis sel kanker mulut oleh caspase-3 jalur dependent caspase-3-pada sel osteosarkoma manusia, yang mengarah ke penghambatan pertumbuhan tumor, fosfatidil inositol 3-kinase dan transformasi neoplastik. Demikian pula dengan myricetin yang juga menunjukkan efek antikanker dengan menginduksi apoptosis dan menghambat proliferasi pada sel leukemia manusia [18]. ...
Article
Kanker merupakan suatu penyakit yang ditandai dengan terjadinya proliferasi sel-sel abnormal yang tidak terkendali. Berdasarkan data dari Global Cancer Statistics (GLOBOCAN) menunjukkan bahwa tahun 2020 terdapat jumlah kasus kanker sebanyak 19,3 juta dan kasus kematian akibat kanker sebanyak 10 juta di dunia. Penggunaan obat tradisional di Indonesia mulai berkembang pesat dan digunakan kembali oleh masyarakat sebagai alternatif pengobatan. Buah jamblang (Syzygium cumini L.) merupakan salah satu tumbuhan lokal Indonesia yang termasuk dalam golongan famili Myrtaceae yang memiliki banyak manfaat bagi kesehatan, salah satunya adalah sebagai antikanker. Tujuan dari penulisan ini adalah untuk memberikan informasi dan gambaran mengenai potensi buah jamblang sebagai agen antikanker dan potensinya sebagai bahan pangan fungsional. Metode yang digunakan dalam penelitian ini yaitu Systematic Literature Review (SLR). Penelusuran literatur yakni melalui Pubmed dan Google Scholar menggunakan sumber data elektronik yang terpublikasi dari tahun 2018-2023. Hasil dari pencarian literatur yang memenuhi kriteria inklusi ditemukan sebanyak 36 literatur dari 22.045 temuan. Dari beberapa hasil penelitian menunjukkan bahwa buah jamblang (Syzygium cumini L.) memiliki karakteristik kemopreventif dan kemoterapi. Ekstrak buah dari tanaman tersebut kaya akan antosianin, flavonoid, senyawa fenolik, dan karotenoid yang memiliki aktivitas sebagai antikanker. Selain itu, pada bagian buahnya kaya akan nutrisi penting seperti protein, karbohidrat, serat, dan vitamin yang berguna sebagai sumber nutrasetikal bagi tubuh. Oleh karena itu, buah jamblang (Syzygium cumini L.) berpotensi untuk dikembangkan menjadi bahan pangan fungsional sebagai pendamping terapi pada pasien kanker.
... Similarly, in cancers, flavones have been shown, through polyhydroxylation at positions 3 and 6, to inhibit sarcoplasmic reticulum Ca 2+ -ATPases, thus inhibiting mitochondrial calcium transfer [55]. Furthermore, kaempferol inhibits human osteosarcoma through increases in Ca 2+ with reduced mitochondrial membrane potentials [56]. For WJ9708012, a methoxyflavanone derivative, this anti-cancer pathway has been elucidated as ER-stress dependent increases in GADD153 and GRP78, concomitant with PKC-α-mediated mitochondria stress, resulting in mTOR pathway alterations to cause apoptosis [57]. ...
Article
Full-text available
Hypertension is the leading remediable risk factor for cardiovascular morbidity and mortality in the United States. Excess dietary salt consumption, which is a catalyst of hypertension, initiates an inflammatory cascade via activation of antigen-presenting cells (APCs). This pro-inflammatory response is driven primarily by sodium ions (Na+) transporting into APCs by the epithelial sodium channel (ENaC) and subsequent NADPH oxidase activation, leading to high levels of oxidative stress. Oxidative stress, a well-known catalyst for hypertension-related illness development, disturbs redox homeostasis, which ultimately promotes lipid peroxidation, isolevuglandin production and an inflammatory response. Natural medicinal compounds derived from organic materials that are characterized by their anti-inflammatory, anti-oxidative, and anti-mutagenic properties have recently gained traction amongst the pharmacology community due to their therapeutic effects. Flavonoids, a natural phenolic compound, have these therapeutic benefits and can potentially serve as anti-hypertensives. Flavones are a type of flavonoid that have increased anti-inflammatory effects that may allow them to act as therapeutic agents for hypertension, including diosmetin, which is able to induce significant arterial vasodilation in several different animal models. This review will focus on the activity of flavones to illuminate potential preventative and potential therapeutic mechanisms against hypertension.
... зафиксировали, что флавоноид индуцирует повреждение ДНК и ингибирует экспрессию белка, связанного с репарацией ДНК, в клетках промиелоцитарного лейкоза HL-60 человека [86]. Кемпферол может тормозить развитие остеосаркомы человека в связи с его способностью индуцировать апоптоз раковых клеток через стресс эндоплазматического ретикулума и митохондриально-зависимый путь [87]. Таким образом, анализ данных литературы показывает, что кемпферол, как и кверцетин, проявляет противоопухолевую активность, что позволяет рекомендовать его для комплексной терапии злокачественных заболеваний. ...
Article
Full-text available
Flavonoids are an extensive class of secondary metabolites present in varying concentrations in different parts of plants. Medicinal raw materials containing flavonoids have been used in traditional medicine in various countries for centuries and are also used in modern medicine for the production of drugs. Compared to other groups of secondary metabolites, flavonoids are often present in relatively large amounts. Interest in flavonoids is driven by constantly updated data on their biological activity and their wide distribution in the plant world. This review focuses on the two most commonly occurring polyphenolic compounds in plants, quercetin and kaempferol. The article describes the main glycosidic forms of the flavonoids under consideration and the modern results of studying their biological activity, namely their antioxidant, anti-allergic, anti-inflammatory, cardioprotective, and anti-tumor properties. In addition, some of the mechanisms for implementing the above types of biological activity are discussed. The analysis suggests that further in-depth pharmacological research on these flavonoids and the development of new advanced drugs based on them is promising. The cores of quercetin and kaempferol can be recommended for chemical modification to obtain highly active compounds with antioxidant, anti-allergic, anti-inflammatory, cardioprotective, and anti-tumor activities. Keywords: flavonoids, quercetin, kaempferol, glycosidic forms, biological activity, antioxidant activity, anti-allergic activity, anti-inflammatory activity, cardioprotective activity, anti-tumor activity, mechanisms of action, drugs, chemical modification.
Article
Full-text available
Kaempferol is a flavonoid present in many eatable plants. Researchers discovered a link between consuming foods high in kaempferol and lowering the threat for acquiring many diseases, including cardiovascular disease, diabetes, obesity and cancer. Kaempferol can inhibit AHR transcription, modulate ERK signalling pathway and NF-κB pathways, block MAPK and AP1 signalling pathways and perform more anticancer roles. Kaempferol also has the ability to act against diabetes via suppressed phosphorylation of (IKK) IkB kinase, (IRS-1) insulin receptor substrate-1 and (IKK) IkB kinase through the hepatic IKK/ NF-B signalling pathways and significantly enhanced insulin secretion and synthesis. Kaempferol protects cardiomyocytes from anoxia/reoxygenation (A/R)-induced damage by lowering LDH release, improving cell survival, reducing A/R-induced ROS formation, and release of cytochrome c. This knowledge may aid in understanding health advantages of medicinal plants that contain kaempferol and may lead to the development of the flavonoid as a potential agents for disease aversion and therapy
Article
Irinotecan (IRI), an anticancer drug to treat colon cancer patients, causes cytotoxic effects on normal cells. Phenethyl isothiocyanate (PEITC), rich in common cruciferous plants, has anticancer activities (induction of cell apoptosis) in many human cancer cells, including colon cancer cells. However, the anticancer effects of IRI combined with PEITC on human colon cancer cells in vitro were unavailable. Herein, the aim of this study is to focus on the apoptotic effects of the combination of IRI and PEITC on human colon cancer HCT 116 cells in vitro. Propidium iodide (PI) exclusion and Annexin V/PI staining assays showed that IRI combined with PEITC decreased viable cell number and induced higher cell apoptosis than that of IRI or PEITC only in HCT 116 cells. Moreover, combined treatment induced higher levels of reactive oxygen species (ROS) and Ca ²⁺ than that of IRI or PEITC only. Cells pre‐treated with N ‐acetyl‐ l ‐cysteine (scavenger of ROS) and then treated with IRI, PEITC, or IRI combined with PEITC showed increased viable cell numbers than that of IRI or PEITC only. IRI combined with PEITC increased higher caspase‐3, ‐8, and ‐9 activities than that of IRI or PEITC only by flow cytometer assay. IRI combined with PEITC induced higher levels of ER stress‐, mitochondria‐, and caspase‐associated proteins than that of IRI or PEITC treatment only in HCT 116 cells. Based on these observations, PEITC potentiates IRI anticancer activity by promoting cell apoptosis in the human colon HCT 116 cells. Thus, PEITC may be a potential enhancer for IRI in humans as an anticolon cancer drug in the future.
Preprint
Full-text available
Background Osteosarcoma (OS) is the most common primary bone sarcoma. OS is most likely to occur in adolescents. Based on clinical experience, Huayan Capsules (HYCA) has adjuvant therapeutic effects in OS patients. Through network pharmacology, molecular docking and cell experiments, we sought to investigate the active components, targets and mechanism of HYCA in the treatment of OS. Methods The active components and targets of HYCA were found using the TCMSP and TCMID. GeneCards, TTD, and OMIM were used to find OS-related targets. The KEGG and GO enrichment were used to study PPI. Using Auto Dock Vina, the substance was molecularly docked with proteins related to OS. Finally, cell experiments were carried out to support the above conclusions. Results It was found HYCA had 1703 targets and 239 active molecules. Between OS and HYCA, there were 220 intersection targets. The PPI network revealed TP53, AKT1 were among the 25 primary targets of HYCA. GO enrichment revealed the genes were enriched in cellular reactions to hormones and other substances. KEGG enrichment revealed the genes were enriched in 196 pathways, mostly related to cancer, such as the PI3K–AKT and MAPK signaling pathways. According to molecular docking, quercetin, kaempferol, and beta-sitosterol have strong binding abilities with AKT1 and TP53. Cell experiments showed beta-sitosterol could inhibit the growth and wound healing formation of OS cells and promote apoptosis. Conclusions We predict the active compounds and potential targets of HYCA. Beta-sitosterol, one of the leading monomers of HYCA, can inhibit proliferation, migration of OS cells and induce apoptosis.
Article
This study sought to determine the anticancer effect of kaempferol, a glycone-type flavonoid glycoside with various pharmacological benefits, on human oral cancer MC-3 cells. In vitro studies comprised a 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide assay, annexin V and propidium iodide staining, western blotting analysis, and acridine orange staining, while the in vivo studies entailed a xenograft model, hematoxylin and eosin staining, and TdT-mediated dUTP-biotin nick end labelling. In vitro, kaempferol reduced the rate of survival of MC-3 cells, mediated intrinsic apoptosis, increased the number of acidic vesicular organelles, and altered the expression of autophagy-related proteins. Further, treatment with the autophagy inhibitors revealed that the induced autophagy had a cytoprotective effect on apoptosis in kaempferol-treated MC-3 cells. Kaempferol also decreased the expression of phosphorylated extracellular signal-regulated kinase and increased that of phosphorylated c-Jun N-terminal kinase (p-JNK) and phosphorylated p38 kinase in MC-3 cells, suggesting the occurrence of mitogen-activated protein kinase-mediated apoptosis and JNK-mediated autophagy. In vivo, kaempferol reduced tumor growth inducing apoptosis and autophagy. These results showed that kaempferol has the potential use as an adjunctive agent in treating oral cancer.
Article
Full-text available
Insulin-like growth factor-I receptor (IGF-IR) is an important mediator of tumor cell survival and shows prognostic significance in sarcoma. To explore potential therapeutic strategies for interrupting signaling through this pathway, we assessed the ability of cyclolignan picropodophyllin (PPP), a member of the cyclolignan family, to selectively inhibit the receptor tyrosine kinase activity of IGF-IR in several sarcoma cell line model systems. Of the diverse sarcoma subtypes studied, osteosarcoma cell lines were found to be particularly sensitive to IGF-IR inhibition, including several multidrug resistant osteosarcoma cell lines with documented resistance to various conventional anticancer drugs. PPP shows relatively little toxicity in human osteoblast cell lines when compared with osteosarcoma cell lines. These studies show that PPP significantly inhibits IGF-IR expression and activation in both chemotherapy-sensitive and chemotherapy-resistant osteosarcoma cell lines. This inhibition of the IGF-IR pathway correlates with suppression of proliferation of osteosarcoma cell lines and with apoptosis induction as measured by monitoring of poly(ADP-ribose) polymerase and its cleavage product and by quantitative measurement of apoptosis-associated CK18Asp396. Importantly, PPP increases the cytotoxic effects of doxorubicin in doxorubicin-resistant osteosarcoma cell lines U-2OS(MR) and KHOS(MR). Furthermore, small interfering RNA down-regulation of IGF-IR expression in drug-resistant cell lines also caused resensitization to doxorubicin. Our data suggest that inhibition of IGF-IR with PPP offers a novel and selective therapeutic strategy for ostosarcoma, and at the same time, PPP is effective at reversing the drug-resistant phenotype in osteosarcoma cell lines.
Article
We have investigated whether Ginkgo biloba extract (EGb 761) induces apoptosis of oral cavity cancer cells and attempted to characterize the apoptotic pathway activated by EGb 761. The inhibition of SCC 1483 oral cavity cancer cells proliferation was noted from 250 mu g/ml of EGb 761. Apoptosis was observed after 24h of incubation with 250 mu g/ml EGb 761 and occurred in a time- and dose-dependent manner. Apoptosis was confirmed by DNA fragmentation and PARP cleavage. Co-treatment with the caspase inhibitor (z-VAD-fmk) inhibited apoptosis and PARP cleavage induced by EGb 761. Caspase-3 activity was upregulated by EGb 761 but reduced to the control level by co-treating with z-VAD-fmk. In summary, EGb 761 induces apoptosis of oral cavity cancer cells and caspase-3 is activated in this apoptosis. Therefore, EGb 761 may be considered as a possible chemopreventive agent against oral cavity cancer. (c) 2004 Elsevier Ltd. All rights reserved.
Article
The relationship between the structure of 42 flavonoids and their antioxidant and antiradical activities was elucidated by heat-induced oxidation in a beta-carotene and linoleic acid system and by the 1,1-diphenyl-2-picrylhydrazyl decoloration test. From seven structurally divergent groups of flavonoids, only flavonols with a free hydroxyl group at the C-3 position of the flavonoid skeleton showed high inhibitory activity to beta-carotene oxidation. Antiradical activity depended on the presence of a flavonol structure or free hydroxyl group at the C-4' position. The effect of the 4'-hydroxyl was strongly modified by other structural features, such as the presence of free hydroxyls at C-3 and/or C-3' and a C2-C3 double bond.
Article
In this study, we investigated the effects of DADS on human colon cancer cell line COLO 205 on cell cycle arrest and apoptosis in vitro. After 24 h treatment of COLO 205 cells with DADS, the dose- and time-dependent decreases of viable cells were observed and the IC50 was 22.47 μM. The decreased percentages of viable cells are associated with the production of ROS. Treatment of COLO 205 cells with DADS resulted in G2/M phase arrest and apoptosis occurrence through the mitochondrial-pathway (Bcl-2, Bcl-xL down-regulation and Bak, Bax up-regulation). DADS increased cyclin B, cdc25c-ser-216-9 and Wee1 but did not affect CDK1 protein and gene expression within 24 h of treatment. DADS-induced apoptosis was examined and confirmed by DAPI staining and DNA fragmentation assay. DADS promoted caspase-3, -8 and -9 activity and induced apoptosis were accompanied by increasing the levels of Fas, phospho-Ask1 and -JNK, p53 and decreasing the mitochondrial membrane potential which then led to release the cytochrome c, cleavage of pro-caspase-9 and -3. The COLO 205 cells were pre-treated with JNK inhibitor before leading to decrease the percentage of apoptosis which was induced by DADS. Inhibition of caspase-3 activation blocked DADS-induced apoptosis on COLO 205 cells.
Article
CumuIative and excessive exposure to estrogens is associated with increased breast cancer risk. The traditional mechanism explaining this association is that estrogens affect the rate of cell division and apoptosis and thus manifest their effect on the risk of breast cancer by affecting the growth of breast epithelial tissues. Highly proliferative cells are susceptible to genetic errors during DNA replication. The action of estrogen metabolites offers a complementary genotoxic pathway mediated by the generation of reactive estrogen quinone metabolites that can form adducts with DNA and generate reactive oxygen species through redox cycling. In this chapter, we discussed a novel mitochondrial pathway mediated by estrogens and their cognate estrogen receptors (ERs) and its potential implications in estrogen-dependent carcinogenesis. Several lines of evidence are presented to show: (1) mitochondrial localization of ERs in human breast cancer cells and other cell types; (2) a functional role for the mitochondrial ERs in regulation of the mitochondrial respiratory chain (MRC) proteins and (3) potential implications of the mitochondrial ER-mediated pathway in stimulation of cell proliferation, inhibition of apoptosis and oxidative damage to mitochondrial DNA. The possible involvement of estrogens and ERs in deregulation of mitochondrial bioenergetics, an important hallmark of cancer cells, is also described. An evolutionary view is presented to suggest that persistent stimulation by estrogens through ER signaling pathways of MRC proteins and energy metabolic pathways leads to the alterations in mitochondrial bioenergetics and contributes to the development of estrogen-related cancers.
Article
ITR-284, a potent anti-leukemia agent of carboxamide derivative, has been shown to inhibit the proliferation of leukemia cells. In this study, the underlying molecular mechanisms in vitro and anti-leukemia activity in vivo of ITR-284 were investigated. ITR-284 reduced the cell viability and induced apoptosis in HL-60 and WEHI-3 leukemia cells. Following exposure of cells to 30 nM of ITR-284, there is a time-dependent decrease in the mitochondrial membrane potential (DeltaPsi(m)) and an increase in the reactive oxygen species (ROS). ITR-284 treatment also caused a time-dependent increase of Fas/CD95, cytosolic cytochrome c, cytosolic active form of caspase-8/-9/-3, cytosolic Apaf-1 and Bax, and the decrease of Bcl-2. However, the ITR-284-induced caspase-8/-9 and -3 activities can be blocked by pan-caspase inhibitor (Z-VAD-FMK). In addition, the anti-leukemia effects of ITR-284 in vivo were further evaluated in BALB/c mice inoculated with WEHI-3 cells. Orally treatment with ITR-284 (2 and 10mg/kg/alternate day for 7 times) increased the survival rate and prevented the loss of body weight in leukemia mice. The enlargement of spleen and infiltration of immature myeloblastic cells into spleen red pulp were significantly reduced in ITR-284-treated mice compared with control mice. Moreover, ITR-284 application can enhance the anti-leukemia effect of all-trans retinoic acid (ATRA). These results revealed that ITR-284 acted against both HL-60 and WEHI-3 in vitrovia both intrinsic and extrinsic apoptotic signaling pathways, and exhibited an anti-leukemic effect in a WEHI-3 orthotopic mice model of leukemia.
Article
It is estimated that 75-85% of all chronic diseases are linked to lifestyle-related and environmental factors. The development of colon cancer is positively associated with obesity and inversely associated with the intake of dietary fibre, fruit and vegetable. Apple juice is the most widely consumed fruit beverage in Germany. It contains a specific spectrum of polyphenols and other components that may reduce the risk of colon cancer. Epidemiologic studies suggest an inverse correlation between apple consumption and colon cancer risk, although the mechanisms for these observations are not clear. The present review summarizes the preventive potential of apple juices and different apple constituents on biomarkers related to colon carcinogenesis with special focus on the in vivo evidence and the cancer promoting condition of obesity. However, under the cancer promoting condition of obesity, apple juice did not show cancer-preventive bioactivity. In our experiments a cancer-preventive bioactivity of apple juice is lacking in rats under the cancer-promoting condition of obesity. To further investigate, whether this lack of efficacy observed in obese rats might be representative for obese individuals human intervention studies on high risk groups such as obese or diabetic individuals are of interest and will be conducted.
Article
Capsaicin, a pungent compound found in hot chili peppers, induces apoptotic cell death in various cell lines, however, the precise apoptosis signaling pathway is unknown. Here, we investigated capsaicin-induced apoptotic signaling in the human breast cell line MCF10A and found that it involves both endoplasmic reticulum (ER) stress and calpain activation. Capsaicin inhibited growth in a dose-dependent manner and induced apoptotic nuclear changes in MCF10A cells. Capsaicin also induced degradation of tumor suppressor p53; this effect was enhanced by the ER stressor tunicamycin. The proteasome inhibitor MG132 completely blocked capsaicin-induced p53 degradation and enhanced apoptotic cell death. Capsaicin treatment triggered ER stress by increasing levels of IRE1, GADD153/Chop, GRP78/Bip, and activated caspase-4. It led to an increase in cytosolic Ca(2+), calpain activation, loss of the mitochondrial transmembrane potential, release of mitochondrial cytochrome c, and caspase-9 and -7 activation. Furthermore, capsaicin-induced the mitochondrial apoptotic pathway through calpain-mediated Bid translocation to the mitochondria and nuclear translocation of apoptosis-inducing factor (AIF). Capsaicin-induced caspase-9, Bid cleavage, and AIF translocation were blocked by calpeptin, and BAPTA and calpeptin attenuated calpain activation and Bid cleavage. Thus, both ER stress- and mitochondria-mediated death pathways are involved in capsaicin-induced apoptosis.
Article
In this study, we investigated the effects of 2-(2-fluorophenyl)-6,7-methylenedioxyquinolin-4-one (CHM-1) on cell viability, cell cycle arrest and apoptosis in CT-26 murine colorectal adenocarcinoma cells. For determining cell viability, the MTT assay was used. CHM-1 promoted G2/M arrest by PI staining and flow cytometric analysis. Apoptotic cells were evaluated by DAPI staining. We used CDK1 kinase assay, Western blot analysis and caspase activity assays for examining the CDK1 activity and proteins correlated with apoptosis and cell cycle arrest. The in vivo anti-tumor effects of CHM-1-P were evaluated in BALB/c mice inoculated with CT-26 cells orthotopic model. CHM-1 induced CT-26 cell viability inhibition and morphologic changes in a dose-dependent and time-dependent manner and the approximate IC50 was 742.36 nM. CHM-1 induced significant G2/M arrest and apoptosis in CT-26 cells. CHM-1 inhibited the CDK1 activity and decreased CDK1, Cyclin A, Cyclin B protein levels. CHM-1 induced apoptosis in CT-26 cells and promoted increasing of cytosolic cytochrome c, AIF, Bax, BAD, cleavage of pro-caspase-9, and -3. The significant reduction of caspase-9 and -3 activity and increasing the viable CT-26 cells after pretreated with caspase-9 and -3 inhibitor indicated that CHM-1-induced apoptosis was mainly mediated a mitochondria-dependent pathway. CHM-1-P improved mice survival rate, and enlargement of the spleen and liver metastasis were significantly reduced in groups treated with either 10 mg/kg and 30 mg/kg of CHM-1-P and 5-FU in comparison to these of CT-26/BALB/c mice. Taken together, CHM-1 acted against colorectal adenocarcinoma cells in vitro via G2/M arrest and apoptosis, and CHM-1-P inhibited tumor growth in vivo.
Article
EGb 761, extracted from Ginkgo biloba leaves, has been proven to induce caspase-3-dependent apoptosis in oral cavity cancer cells. Since EGb 761 is a composition of various components, it is important to identify which components are responsible for its anticancer effects to reduce the total dosage and to avoid toxicity. Therefore, the study aimed to determine the effective compounds of EGb 761 that induce apoptosis in oral cavity cancer cells and to identify whether caspase-3 was involved in apoptosis of oral cancer cells by EGb 761 components. The results of cell proliferation assays on oral cavity cancer cells showed that kaempferol and quercetin significantly inhibited cellular proliferation at a concentration of 40 microM. Flow cytometry showed that the antiproliferative effects of each component were due to increased apoptosis. Kaempferol and quercetin induced apoptosis in various oral cancer cell lines (SCC-1483, SCC-25 and SCC-QLL1) and showed cleavage of poly (ADP-ribose) polymerase (PARP). Caspase-3 activity assay revealed that induction of apoptosis by kaempferol and quercetin was caspase-3-dependent. In conclusion, the results suggest that kaempferol and quercetin, two components of EGb 761, effectively induce caspase-3-dependent apoptosis of oral cavity cancer cells and can be considered as possible anti-oral cavity cancer agents.