ArticlePDF Available

A Novel Anticancer Agent, Decursin, Induces G1 Arrest and Apoptosis in Human Prostate Carcinoma Cells

Authors:

Abstract

We isolated a coumarin compound decursin (C(19)H(20)O(5); molecular weight 328) from Korean angelica (Angelica gigas) root and characterized it by spectroscopy. Here, for the first time, we observed that decursin (25-100 micromol/L) treatment for 24 to 96 hours strongly inhibits growth and induces death in human prostate carcinoma DU145, PC-3, and LNCaP cells. Furthermore, we observed that decursinol [where (CH(3))(2)-C=CH-COO- side chain of decursin is substituted with -OH] has much lower effects compared with decursin, suggesting a possible structure-activity relationship. Decursin-induced growth inhibition was associated with a strong G(1) arrest (P < 0.001) in DU145 and LNCaP cells, and G(1), S as well as G(2)-M arrests depending upon doses and treatment times in PC-3 cells. Comparatively, decursin was nontoxic to human prostate epithelial PWR-1E cells and showed only moderate growth inhibition and G(1) arrest. Consistent with G(1) arrest in DU145 cells, decursin strongly increased protein levels of Cip1/p21 but showed a moderate increase in Kip1/p27 with a decrease in cyclin-dependent kinases (CDK); CDK2, CDK4, CDK6, and cyclin D1, and inhibited CDK2, CDK4, CDK6, cyclin D1, and cyclin E kinase activity, and increased binding of CDK inhibitor (CDKI) with CDK. Decursin-caused cell death was associated with an increase in apoptosis (P < 0.05-0.001) and cleaved caspase-9, caspase-3, and poly(ADP-ribose) polymerase; however, pretreatment with all-caspases inhibitor (z-VAD-fmk) only partially reversed decursin-induced apoptosis, suggesting the involvement of both caspase-dependent and caspase-independent pathways. These findings suggest the novel anticancer efficacy of decursin mediated via induction of cell cycle arrest and apoptosis selectively in human prostate carcinoma cells.
2005;65:1035-1044. Cancer Res
Dongsool Yim, Rana P. Singh, Chapla Agarwal, et al.
Apoptosis in Human Prostate Carcinoma Cells
A Novel Anticancer Agent, Decursin, Induces G1 Arrest and
Updated version
http://cancerres.aacrjournals.org/content/65/3/1035
Access the most recent version of this article at:
Cited Articles
http://cancerres.aacrjournals.org/content/65/3/1035.full.html#ref-list-1
This article cites by 38 articles, 9 of which you can access for free at:
Citing articles
http://cancerres.aacrjournals.org/content/65/3/1035.full.html#related-urls
This article has been cited by 8 HighWire-hosted articles. Access the articles at:
E-mail alerts related to this article or journal.Sign up to receive free email-alerts
Subscriptions
Reprints and
.pubs@aacr.orgDepartment at
To order reprints of this article or to subscribe to the journal, contact the AACR Publications
Permissions
.permissions@aacr.orgDepartment at
To request permission to re-use all or part of this article, contact the AACR Publications
Research.
on June 3, 2013. © 2005 American Association for Cancercancerres.aacrjournals.org Downloaded from
A Novel Anticancer Agent, Decursin, Induces G
1
Arrest and
Apoptosis in Human Prostate Carcinoma Cells
Dongsool Yim,1Rana P. Singh,2Chapla Agarwal,2Sookyeon Lee,1Hyungjoon Chi,4
and Rajesh Agarwal2,3
1Department of Pharmacy, Sahm Yook University, Seoul, Korea; 2Department of Pharmaceutical Sciences, School of Pharmacy and
3University of Colorado Cancer Center, University of Colorado Health Sciences Center, Denver, Colorado; and 4Natural Product
Research Institute, Seoul National University, Seoul, Korea
Abstract
We isolated a coumarin compound decursin (C
19
H
20
O
5
;
molecular weight 328) from Korean angelica (Angelica gigas)
root and characterized it by spectroscopy. Here, for the first
time, we observed that decursin (25-100 Mmol/L) treatment for
24 to 96 hours strongly inhibits growth and induces death in
human prostate carcinoma DU145, PC-3, and LNCaP cells.
Furthermore, we observed that decursinol [where (CH
3
)
2
-
C=CH-COO- side chain of decursin is substituted with -OH]
has much lower effects compared with decursin, suggesting a
possible structure-activity relationship. Decursin-induced
growth inhibition was associated with a strong G
1
arrest
(P< 0.001) in DU145 and LNCaP cells, and G
1
, S as well as G
2
-M
arrests depending upon doses and treatment times in PC-3
cells. Comparatively, decursin was nontoxic to human prostate
epithelial PWR-1E cells and showed only moderate growth
inhibition and G
1
arrest. Consistent with G
1
arrest in DU145
cells, decursin strongly increased protein levels of Cip1/p21 but
showed a moderate increase in Kip1/p27 with a decrease in
cyclin-dependent kinases (CDK); CDK2, CDK4, CDK6, and cyclin
D1, and inhibited CDK2, CDK4, CDK6, cyclin D1, and cyclin E
kinase activity, and increased binding of CDK inhibitor (CDKI)
with CDK. Decursin-caused cell death was associated with an
increase in apoptosis (P< 0.05-0.001) and cleaved caspase-9,
caspase-3, and poly(ADP-ribose) polymerase; however, pre-
treatment with all-caspases inhibitor (z-VAD-fmk) only par-
tially reversed decursin-induced apoptosis, suggesting the
involvement of both caspase-dependent and caspase-indepen-
dent pathways. These findings suggest the novel anticancer
efficacy of decursin mediated via induction of cell cycle arrest
and apoptosis selectively in human prostate carcinoma cells.
(Cancer Res 2005; 65(3): 1035-44)
Introduction
Prostate cancer is the most common cancer as well as the second
leading cause of cancer-related deaths in men in Western countries
(1). One out of nine men over 65 years of age is frequently diagnosed
with prostate cancer in the United States (1, 2). Dietary pattern has
been identified as one of the major factors for the difference in
prostate cancer incidence between Western and Asian countries
(2–4). At present, there is no effective therapy available for the
treatment of androgen-independent stage of prostate cancer, which
usually arises after hormonal deprivation/ablation therapy (5).
Cytotoxic chemotherapies or radiotherapy also do not show any
significant improvement in patient condition due to the high
recurrence of apoptosis resistance hormone refractory prostate
cancer, which is responsible for f28,000 deaths per year (1, 6, 7).
The clinical impact of advanced prostate cancer has led to the
exploration of novel treatment modalities as well as anticancer
agents. In this regard, many nutritive and nonnutritive phytochem-
icals with diversified pharmacologic properties have shown
promising responses for the prevention and/or intervention of
various cancers, including prostate cancer (reviewed in refs. 2–4). In
addition, epidemiologic studies, together with extensive basic
laboratory findings, support the potential role of phytochemicals
in the prevention and treatment of prostate cancer (reviewed in
refs. 8–14).
Herbal medicines including conventional and complimentary
medicines are still prevalent, and serve the medicinal needs of a
large population around the world (reviewed in refs. 15, 16),
which could be assessed by the fact that the global herbal
medicine market is currently worth around $30 billion. Herbal
medicine practice can be traced back to the origin of human
culture, having been mentioned in medical protocols such as
traditional Chinese medicine and Indian Ayurvedic herbal
medicine. At present, there is an increased effort for the
isolation of bioactive microchemicals from medicinal plants for
their possible usefulness in the control of various ailments.
Determining molecular structure and mechanisms of action of
bioactive phytochemicals are equally important for providing the
evidence for their efficacy as well as herbal preparations, which
could also potentially lead to the pharmaceutical development of
synthetic or semisynthetic drugs.
Angelica gigas Nakai (Umbelliferae) root has been traditionally
used in Korean folk medicine as a tonic and for treating anemia
and other common diseases (17). There are some reports about
the pharmacologic evaluation of this plant showing antibacterial
and antiamnestic effects, inhibitory effect on acetylcholinesterase,
depression of cardiac contraction, activation of protein kinase C,
and antitumor activity against sarcoma cancer cells (18–21).
Based on the curative potential, efforts have been made to isolate
the active principle from this plant, which led to the isolation of
many coumarin compounds (22). However, there is no report on
the evaluation of anticancer activity of these compounds against
epithelial cancers. In the present investigation, we used the roots
of A. gigas to isolate decursin and decursinol, and for the first
time tested their efficacy against human prostate carcinoma cells.
A structure-activity relationship was observed for these com-
pounds, with decursin showing greater efficacy in inhibiting
growth and causing death of DU145 cells. The anticancer efficacy
of decursin was also evident against human prostate cancer PC-3
Requests for reprints: Rajesh Agarwal, Department of Pharmaceutical Sciences,
School of Pharmacy, University of Colorado Health Sciences Center, 4200 East Ninth
Street, Box C238, Denver, CO 80262. Phone: 303-315-1381; Fax: 303-315-6281; E-mail:
Rajesh.Agarwal@UCHSC.edu.
D2005 American Association for Cancer Research.
www.aacrjournals.org 1035 Cancer Res 2005; 65: (3). February 1, 2005
Research Article
Research.
on June 3, 2013. © 2005 American Association for Cancercancerres.aacrjournals.org Downloaded from
and LNCaP cells. Furthermore, we investigated the mechanistic
rationale for the observed efficacy of decursin, which showed an
arrest in cell cycle progression and apoptosis induction as the
most likely targets in prostate cancer cells.
Materials and Methods
Plant Material, Isolation and Characterization of Decursin and
Decursinol. The roots of A. gigas Nakai ( family Umbelliferae) were verified
by Professor Emeritus Hyungjoon Chi (one of the authors in the present
manuscript). The extraction and fractionation of air-dried powdered root was
done as reported recently (22). Silica gel column chromatography was used to
isolate decursin and decursinol (22). These compounds were characterized by
nuclear magnetic resonance (Bruker AVANCE 400 NMR spectrometer) and
mass spectroscopy (by Jeol JMS-AX505-WA mass spectrometer) at Natural
Product Research Institute, Seoul National University, Seoul, Korea. The
purified coumarin compounds, decursin (C
19
H
20
O
5
) with a molecular weight
of 328 (Fig. 1A), and decursinol (C
14
H
14
O
4
) with a molecular weight of 246
(Fig. 1B), were dissolved in DMSO as stock solutions, and used directly in the
cell culture treatments.
Cell Line and Reagents. Human prostate carcinoma cell lines DU145,
PC-3 and LNCaP, and nonneoplastic human prostate epithelial cell line
PWR-1E were obtained from American Type Culture Collection (Manassas,
VA). Prostate cancer cells were cultured in RPMI 1640 with 10% fetal bovine
serum (Hyclone, Logan, UT) under standard culture conditions (37jC, 95%
humidified air, and 5% CO
2
). PWR-1E cells were cultured in keratinocyte
growth medium supplemented with 10% fetal bovine serum, 5 ng/mL
human recombinant epidermal growth factor and 0.05 mg/mL bovine
pituitary extract (American Type Culture Collection). RPMI 1640 and other
culture materials were from Life Technologies, Inc. (Gaithersburg, MD).
Anti-Cip1/p21 antibody was from Calbiochem (Cambridge, MA), and anti-
Kip1/p27 antibody was from Neomarkers, Inc. (Fremont, CA). Antibodies to
cyclin-dependent kinase (CDK); CDK2, CDK4, and CDK6, cyclin D1 and E,
and Rb-glutathione S-transferase fusion protein were from Santa Cruz
Biotechnology, Inc. (Santa Cruz, CA). Anti-caspases primary antibodies and
anti-cleaved poly(ADP-ribose) polymerase (PARP) antibody were from Cell
Signaling Technology (Beverly, MA). Anti-total PARP antibody was from BD
PharMingen (San Diego, CA). Histone H1 was from Boehringer Mannheim,
Corp. (Indianapolis, IN). [g-
32
P] ATP (specific activity 3,000 Ci/mmol) and
enhanced chemiluminescence (Amersham) detection system were from
Amersham Corp. (Piscataway, NJ). Other chemicals were obtained in their
commercially available highest purity grade.
Cell Growth and Death Assays. Cells were plated at 5,000 cells/cm
2
in
60 mm plates under the standard culture conditions detailed above and
after 24 hours, fed with fresh medium and treated with DMSO control or 25,
50, and 100 Amol/L decursin or decursinol. After the desired treatment time,
cells were collected by a brief trypsinization, and counted in duplicate with
a hemocytometer using Trypan blue dye to score dead cells. Each treatment
and time point had three independent plates. The representative data
shown in this study were reproducible in three independent experiments.
Cell Cycle Analysis by Flow Cytometry. Prostate cancer cells and
PWR-1E cells were grown in 10% serum condition in their respective culture
mediums as mentioned above. At f30% confluency, cells were treated with
DMSO control or 25, 50, and 100 Amol/L decursin and at the end of desired
treatment time, cells were collected and processed for cell cycle analysis.
Briefly, 0.5 10
5
cells were suspended in 0.5 mL of saponin/propidium iodide
solution [0.3% saponin (w/v), 25 Ag/mL propidium iodide (w/v), 0.1 mmol
EDTA, and 10 Ag/mL RNase (w/v) in PBS], and incubated overnight at 4jCin
the dark. Cell cycle distribution was then analyzed by flow cytometry using
fluorescence-activated cell sorting core facility of University of Colorado
Cancer Center. Finally, the percentage of cells in different phases of cell cycle
were determined by ModFit LT cell cycle analysis software.
Immunoblotting and Immunoprecipitation. DU145 cells were grown
in RPMI 1640 medium and treated with DMSO control or 25, 50, and
100 Amol/L decursin for the desired times. Equal volumes of DMSO
(0.1% v/v) were present in each treatment. Following decursin treat-
ments, cell lysates were prepared in nondenaturing lysis buffer [10 mmol
Tris-HCl (pH 7.4), 150 mmol NaCl, 1% Triton X-100, 1 mmol EDTA,
1 mmol EGTA, 0.3 mmol phenylmethylsulfonyl fluoride, 0.2 mmol sodium
orthovanadate, 0.5% NP40, and 5 units/mL aprotinin]. For lysate
preparation, medium was aspirated and cells were washed twice with
ice-cold PBS followed by incubation in lysis buffer for 20 minutes. Then,
cells were scraped and kept on ice for an additional 30 minutes, and
finally cell lysates were cleared by centrifugation at 4jC for 30 minutes
in a tabletop centrifuge. Protein concentration in lysates was determined
using Bio-Rad detergent-compatible protein assay kit (Bio-Rad Labora-
tories, Hercules, CA) by Lowry method.
For Western blotting, 60 to 80 Ag of protein lysate per sample was
denatured in 2SDS-PAGE sample buffer and subjected to SDS-PAGE on
12% or 16% tris-glycine gel as published recently (23). The separated
proteins were transferred on to nitrocellulose membrane followed by
blocking of membrane with 5% nonfat milk powder (w/v) in TBS
(10 mmol Tris, 100 mmol NaCl, 0.1% Tween 20) for 1 hour at room
temperature or overnight at 4jC. Membranes were probed for the
protein levels of Cip1/p21, Kip1/p27, CDK2, CDK4, CDK6, cyclin D1, and
E using specific primary antibodies followed by peroxidase-conjugated
appropriate secondary antibody, and visualized by enhanced chemilumi-
nescence (Amersham) detection system. Similarly, for apoptotic mole-
cules, cleaved caspase-9, cleaved caspase-3, total PARP and cleaved PARP
were probed using their specific primary antibodies followed by
appropriate secondary antibody and enhanced chemiluminescence
visualization. Membranes were stripped and reprobed with h-actin
primary antibody as a protein loading control.
For CDK inhibitor (CDKI)-CDK binding study, DU145 cells were
treated with 100 Amol/L decursin for 24 hours and whole cell lysates
were prepared. For each sample, 200 Ag protein lysates was cleared with
protein A/G-plus agarose beads (Santa Cruz) for 45 minutes at 4jC.
Cip1/p21 and Kip1/p27 were then immunoprecipitated from protein
lysates using specific antibody (2 Ag) incubation for 6 hours followed by
addition of 25 AL of protein A/G-plus agarose beads and rocking
overnight at 4jC. Immunoprecipitates were washed thrice with lysis
buffer, and samples were boiled in 2sample buffer for 5 minutes
followed by centrifugation. The resulting clear supernatants were
subjected to SDS-PAGE on 16% gel and Western blotting. Membranes
were then probed with and visualized for CDK2, CDK4, and CDK6 as
detailed above.
Kinase Assays. CDK2 and cyclin E-associated histone H1 kinase
activity was determined as described recently (23). Briefly, 200 Agof
protein lysates from each sample was precleared with protein A/G-plus
agarose beads, and CDK2 and cyclin E proteins were immunoprecipi-
tated using anti-CDK2 and anti–cyclin E antibodies (2 Ag) and protein
A/G-plus agarose beads, respectively. Beads were washed thrice with lysis
buffer, and finally with kinase assay buffer [50 mmol Tris-HCl (pH 7.4),
10 mmol MgCl
2
and 1 mmol DTT]. Phosphorylation of histone H1 was
measured by incubating the beads with 30 AL of ‘‘hot’’ kinase solution
[0.25 AL (2.5 Ag) of histone H1, 0.5 AL(5ACi) of g-
32
P-ATP, 0.5 ALof
0.1 mmol ATP, and 28.75 AL of kinase buffer] for 30 minutes at 37jC.
The reaction was stopped by boiling the samples in SDS sample buffer
for 5 minutes. The samples were then subjected to 12% SDS-PAGE, and
the gel was dried and analyzed by autoradiography (23). Similarly, to
determine CDK4, CDK6, and cyclin D1-associated Rb kinase activities,
these proteins were immunoprecipitated using specific antibodies, and
beads conjugated with antibody and proteins were washed thrice with
Rb-lysis buffer [50 mmol HEPES-KOH (pH 7.5), 150 mmol NaCl, 1 mmol
EDTA, 2.5 mmol EGTA, 1 mmol DTT, 80 mmol h-glycerophosphate,
1 mmol NaF, 0.1 mmol sodium orthovanadate, 0.1% Tween 20, 10%
glycerol, 1 mmol phenylmethylsulfonyl fluoride, and 10 Ag/mL leupeptin and
aprotonin] and twice with Rb-kinase assay buffer [50 mmol HEPES-KOH
(pH 7.5), 2.5 mmol EGTA, 10 mmol h-glycerophosphate, 1 mmol NaF, 10
mmol MgCl
2
, 0.1 mmol sodium orthovanadate, and 1 mmol DTT].
Phosphorylation of Rb was measured by incubating the beads with 30 AL
of hot Rb-kinase solution (2 Ag of Rb-glutathione S-transferase fusion
Cancer Research
Cancer Res 2005; 65: (3). February 1, 2005 1036 www.aacrjournals.org
Research.
on June 3, 2013. © 2005 American Association for Cancercancerres.aacrjournals.org Downloaded from
Figure 1. Structure, growth inhibitory, and cell death effects of decursin and decursinol on DU145 cells. Chemical structure of decursin (A) and decursinol (B),
isolated from the roots of A. gigas . To assess the effect of decursin and decursinol on exponentially growing DU145 cells, 5,000 cells/cm
2
were plated in 60 mm dishes
and the next day, cells were treated either DMSO vehicle control or 25, 50, and 100 Amol/L doses of decursin or decursinol in complete medium. After 24, 48, 72,
and 96 hours of these treatments, total cells were counted using a hemocytometer (C), dead cells were scored using Trypan blue dye exclusion method (D).
Mean FSE of three independent plates; each sample was counted in duplicate. DN, decursin; DL, decursinol; *, P < 0.05;
$
,P< 0.01;
#
,P< 0.001 versus control.
A Novel Anticancer Agent Decursin against Prostate Carcinoma Cells
www.aacrjournals.org 1037 Cancer Res 2005; 65: (3). February 1, 2005
Research.
on June 3, 2013. © 2005 American Association for Cancercancerres.aacrjournals.org Downloaded from
protein, 5 ACi of g-
32
P-ATP, and 0.1 mmol ATP in Rb-kinase buffer) for
30 minutes at 37jC. Reaction was stopped by boiling the samples in
5SDS sample buffer for 5 minutes. Samples were analyzed by SDS-PAGE,
and the gel was dried and subjected to autoradiography (23).
Apoptotic Cell Death Assay. To quantify decursin-induced apoptotic
death of human prostate carcinoma DU145 cells, annexin V, and propidium
iodide staining was done followed by flow cytometry, as described recently
(23, 24). Briefly, after treatment (DMSO control, 50 and 100 Amol/L decursin
for 24 and 48 hours), cells were collected and subjected to annexin V and
propidium iodide staining using Vybrant Apoptosis Assay Kit2 (Molecular
Probes, Inc., Eugene, OR) and following the step-by-step protocol provided
by the manufacturer. Annexin V binds to phospatidylserine, which is
translocated from inner to outer leaflet of the plasma membrane in
apoptotic cells. In all-caspases inhibitor (z-VAD-fmk; Enzyme Systems
Products, ICN Pharmaceuticals, Inc., Aurora, OH) treatment, cells were
pretreated with the inhibitor for 2 hours before decursin treatment.
Caspase Activity Assay. Caspase-3 activity was assayed by colori-
metric protease assay ApoTarget kit (BioSource International, Inc., CA)
following manufacturer’s protocol as published recently (24). Briefly, at the
end of treatment with either 100 Amol/L dose of decursin, or all-caspases
inhibitor z-VAD-fmk (100 Amol/L), or both, in which cells were pretreated
for 2 hours with the inhibitor for a total of 48 hours, cells were collected and
cell lysates prepared in cell lysis buffer (TBS containing detergent). Protein
lysate (100 Ag per sample) was mixed with 2reaction buffer and 200 Amol/
L substrate (DEVD-pNA for caspase-3) and incubated at 37jC for 3 hours in
the dark. These assay conditions are standardized and products of the
reaction remain in the linear range of detection. Developed color was
measured at 405 nm in microplate reader, and blank reading was subtracted
from each sample reading before calculation.
Statistical Analysis. The data were analyzed using the Jandel Scientific
SigmaStat 2.03 software. Student’s ttest was employed to assess the
statistical significance of difference between control and decursin-treated or
decursinol-treated groups. A statistically significant difference was consid-
ered to be present at P< 0.05. Autoradiograms/bands were scanned with
Adobe Photoshop 6.0 (Adobe Systems, Inc. San Jose, CA).
Results
Nuclear Magnetic Resonance and Mass Characterization of
Decursin and Decursinol. Decursin: white crystals from CH
3
OH.
[
1
H] NMR (400 MHz, CDCl
3
)y
H
(ppm): 7.57 (1H, d, J = 9.5 Hz, H-4),
7.14 (1H, s, H-5), 6.78 (1H, s, H-8), 6.21 (1H, d, J = 9.5 Hz, H-3), 5.64
(1H, s, H-2V), 5.06 (1H, t, J = 4.6 Hz, H-3V), 3.17 (1H, dd, J = 16.0, 4.2
Hz, H-4V
a
), 2.85 (1H, dd, J = 16.0, 4.2 Hz, H-4V
b
), 2.12 (3H, s, 3V-CH
3
),
1.86 (3H, s, H-400), 1.36 (3H, s, gem-CH
3
), 1.34 (3H, s, gem-CH
3
). Fast
atom bombardment mass spectroscopy: m/z(rel. int., %): 329 (100)
[M + H]
+
, 228 (100), 213 (95), 154 (45), 136 (45; Fig. 1A). Decursinol:
white crystals from CH
3
OH. [
1
H] NMR (400 MHz, CDCl
3
)
H
(ppm):
7.57 (1H, d, J = 9.5 Hz, H-4), 7.17 (1H, s, H-5), 6.77 (1H, s, H-8), 6.21
(1 H, d, J = 9.2 Hz, H-3), 3.86 (1H, t, J = 4.8 Hz, H-3V), 3.10 (1H, dd,
J = 17.0, 4.5 Hz, H-4V
a
) 2.83 (1H, dd, J = 17.0, 4.5 Hz, H-4V
b
) 1.39 (3H, s,
gem-CH
3
), 1.36 (3H, s, gem-CH
3
). Fast atom bombardment mass
spectroscopy: m/z(rel. int., %): 247 (100) [M + H]
+
, 246 (85), 154
(100), 136 (90; Fig. 1B).
Effects of Decursin and Decursinol on Growth and Death of
DU145 Cells. To assess the biological activity of these com-
pounds in terms of cell growth and death, DU145 cells were
treated with 25, 50, and 100 Amol/L doses of decursin and
decursinol for 24, 48, 72, and 96 hours. Decursin (25-100 Amol/L)
showed a strong dose- and time-dependent inhibition of cell
growth, accounting for 22% to 51% (P< 0.05-0.001), 21% to 68%
(P< 0.01-0.001), 9% to 72% (P< 0.001), and 42% to 90%
(P< 0.001) growth inhibition after 24, 48, 72, and 96 hours of
treatment, respectively (Fig. 1C). However, similar treatment with
decursinol (25-50 Amol/L for 24, 48, 72, and 96 hours) showed
almost half the efficacy of decursin, accounting for 11% to 36%
(P< 0.05-0.001), 9% to 45% (P< 0.01-0.001), 9% to 27% (P< 0.001),
and 31% to 51% (P< 0.001) growth inhibition, respectively
(Fig. 1C). These data suggest that decursin is much more potent
in inhibiting the growth of DU145 as compared with decursinol.
The differences in biological activity of these two compounds
could be attributed to the difference in their chemical structures
(Fig. 1Aand B). The increased activity of decursin could be, most
likely, due to (CH
3
)
2
-C=CH-COO- side chain of decursin which
is substituted with OH group in decursinol (Fig. 1Aand B).
Furthermore, we observed that both compounds have cytotoxic
effect on DU145 cells, in which similar treatment (25-100 Amol/L
for 24, 48, 72, and 96 hours) with decursin caused 15% to 45%
(P< 0.05-0.001) cell death versus 11% to 12% in controls (Fig. 1D),
whereas decursinol caused 13% to 32% (P< 0.05-0.001) cell death
as compared with 8% to 13% in controls (Fig. 1D). When these
cytotoxicity data were compared between these two agents, at
similar doses and treatment times, decursin showed a relatively
better cell death effect compared with decursinol, except at
48 hours of treatment; however, it was statistically significant
(decursin versus decursinol) only at 25 Amol/L dose for 96-hour
(P< 0.05) and 100 Amol/L dose for 72-hour (P< 0.001)
treatments. These data suggest a relatively higher cytotoxic effect
of decursin as compared with decursinol, and further support a
possible structure-activity relationship in their overall effects in
DU145 cells. Next, we investigated the growth inhibitory and
cytotoxic effects of decursin on human prostate cancer PC-3 and
LNCaP cells, and human prostate epithelial PWR-1E cells, and
mechanisms associated with these effects of decursin in DU145
cells.
Effect of Decursin on Growth and Death of Human
Prostate Cancer PC-3 and LNCaP Cells, and Human Prostate
Epithelial PWR-1E Cells. As compared with DU145 cells,
decursin (25-100 Amol/L) treatment of PC-3 cells for 24, 48, 72,
and 96 hours, showed stronger dose- and time-dependent growth
inhibitory effect accounting for 7% to 54% (P< 0.05-0.001), 25% to
75% (P< 0.05-0.001), 36% to 84% (P< 0.05-0.001), and 49% to 91%
(P< 0.05-0.001) inhibition, respectively (Fig. 2A). However, cytotoxic
effect of decursin in PC-3 cells was limited to higher doses only,
showing 3% to 18% (P> 0.05 to < 0.05-0.001) dead cells as compared
with 2% to 5% in controls (Fig. 2A). The growth inhibitory effect of
decursin in LNCaP cells was stronger than both DU145 and PC-3
cells. In similar decursin treatment for 24, 48, 72, and 96 hours, 9% to
72% (P> 0.05 to < 0.005-0.001), 48% to 83% (P< 0.005-0.001), 76% to
94% (P< 0.001), and 79% to 98% (P< 0.001) cell growth inhibition was
observed in LNCaP cells, respectively (Fig. 2B). The cytotoxic effect
of decursin in LNCaP cells was also greater than that of DU145 and
PC-3 cells accounting for 16% to 52% (P> 0.05 to < 0.05-0.001) cell
death as compared with 8% to 16% in controls (Fig. 2B). Overall,
these results suggested growth inhibitory and cytotoxic effects of
decursin in human prostate carcinoma PC-3, LNCaP as well as
DU145 cells.
Furthermore, using nonneoplastic human prostate epithelial
PWR-1E cells, we assessed whether decursin has any differential
sensitivity to normal versus cancer cells. Surprisingly, we did not
observe any increase in dead cells following decursin treatment;
it was even significantly lower (P< 0.05 versus control) following
72 hours of treatment at the 100 Amol/L dose (Fig. 2C). A
decrease in cell number, however, was observed after decursin
treatment of PWR-1E cells, but it was not as strong as in the
case of prostate carcinoma cells (Fig. 2C). For instance, growth
Cancer Research
Cancer Res 2005; 65: (3). February 1, 2005 1038 www.aacrjournals.org
Research.
on June 3, 2013. © 2005 American Association for Cancercancerres.aacrjournals.org Downloaded from
inhibition caused by 50 to 100 Amol/L doses of decursin was
56% to 75% (48 hours) and 64% to 84% (72 hours) in PC-3 cells;
and 58% to 83% (48 hours) and 77% to 94% (72 hours) in LNCaP
cells as compared with 14% to 62% (48 hours) and 41% to 77%
(72 hours) in PWR-1E cells (Fig. 2). After 24 hours of decursin
treatment, we did not observe any considerable growth
inhibition or cell death in PWR-1E cells (data not shown).
These results suggest that decursin is not toxic to prostate
epithelial cells; however, it inhibits their growth with lesser
efficacy as compared with cancer cells in culture.
Decursin Induces a Strong G
1
Arrest in Cell Cycle
Progression of Prostate Cancer Cells. Inhibition of deregulated
cell cycle progression in cancer cells is an effective strategy to
halt tumor growth (14, 25). Because we observed a strong growth
Figure 2. Growth inhibitory and cell death effects of decursin on PC-3, LNCaP, and PWR-1E cells. To assess the effect of decursin on exponentially growing
human PCA PC-3 (A), LNCaP (B) cells, and nonneoplastic prostate epithelial PWR-1E cells (C), 5,000 cells/cm
2
were plated in 60 mm dishes and the next day, cells
were treated either DMSO vehicle control or 25, 50, and 100 Amol/L doses of decursin in complete medium. After the indicated treatment times, total cells were
counted using a hemocytometer, and dead cells were scored using Trypan blue dye exclusion method. Mean FSE of three independent plates; each sample was
counted in duplicate. DN, decursin; DL, decursinol; *, P < 0.05;
$
,P< 0.01;
#
,P< 0.001 versus control.
A Novel Anticancer Agent Decursin against Prostate Carcinoma Cells
www.aacrjournals.org 1039 Cancer Res 2005; 65: (3). February 1, 2005
Research.
on June 3, 2013. © 2005 American Association for Cancercancerres.aacrjournals.org Downloaded from
inhibitory effect of decursin, we then analyzed its possible
inhibitory effect on cell cycle progression following 25, 50, and
100 Amol/L doses of decursin treatment for 12 to 96 hours. We
observed that decursin induces G
1
arrest, which started as early
as 12 hours after the treatment in DU145 cells (data not shown).
An optimum dose-dependent effect was observed at 24 hours of
treatment where decursin caused 52%, 65%, and 78% DU145 cells
in G
1
phase (P< 0.001) as compared with 45% in control (Fig. 3A
and B); this effect started diminishing after 48 hours of treatment
(data not shown). An increase in G
1
cell population was mostly
at the expense of S phase cells (P< 0.05-0.01) with a minimal
decrease in G
2
-M cell population (Fig. 3B).
Similarly, in PC-3 cells, a strong G
1
arrest was observed at 25 and
50 Amol/L decursin treatment for 24 hours (Fig. 4A), which started
diminishing at later treatment times leading to a moderate G
2
-M
arrest (data not shown). However, 100 Amol/L dose of decursin
invariably showed an increase in S and/or G
2
-M phase cell
population in PC-3 cells at all treatment times (Fig. 4A; data for
later treatment times not shown). Similar to DU145 and PC-3 cells,
decursin (25-50 Amol/L) caused significant increase in G
1
cell
population at the expense of S as well as G
2
-M phase cell
populations in LNCaP cells (Fig. 4B). However, in contrast to
DU145 and PC-3 cells, this effect of decursin in LNCaP cells did
not diminish and remained sustained even at 96 hours of the
treatment (data not shown). Higher doses (100 Amol/L) of
decursin, interestingly, showed a time-dependent increase in G
1
arrest in LNCaP cells (data not shown). Furthermore, in
nonneoplastic prostate epithelial PWR-1E cells, only a moderate
increase in G
1
arrest was observed only at higher doses (100
Amol/L) of decursin treatment (Fig. 4C). However, both lower
doses (25 and 50 Amol/L) that were effective in prostate cancer
cells did not show any cell cycle effect in PWR-1E cells (Fig. 4C;
and data not shown). These results suggest that inhibition of
deregulated cell cycle progression could be one of the molecular
events associated with selective anticancer efficacy of decursin in
prostate cancer cells.
Decursin Up-regulates Cip1/p21 and Kip1/p27 Levels, and
Inhibits CDK- and Cyclin-Associated Kinase Activity. Since,
we observed optimum G
1
cell cycle arrest by decursin at 24 hours
in DU145 cells, and a relatively similar effect at 48 hours (data
not shown), total cell lysates were prepared following decursin
treatment of cells at 25, 50, and 100 Amol/L doses for 24 and
48 hours. Western blot analysis of total cell lysates (80 Ag/sample)
showed a strong dose-dependent increase in the expression of
Cip1/p21, which was moderate in case of Kip1/p27 (Fig. 5A).
Reprobing of membranes for h-actin confirmed equal protein
loading (Fig. 5A,bottom). In the studies analyzing its effect on
the levels of CDKs and cyclins associated with G
1
phase,
decursin treatment for 24 hours did not show any alteration in
protein levels of CDK2, CDK4, CDK6, cyclin D1, and cyclin E;
however, 48 hours of treatment showed a dose-dependent
decrease in the expression of these proteins except cyclin E
(Fig. 5B).
Next, cell lysates were assayed for CDK2, CDK4, CDK6, cyclin D1,
and cyclin E kinase activity. As shown in Fig. 5C, compared with
control, decursin treatment resulted in almost complete inhibition
in CDK2, cyclin E, and cyclin D1 kinase activity; however, the
inhibitory effect of decursin on CDK4- and CDK6-associated kinase
activity was of lower magnitude (Fig. 5C). We anticipated that the
observed inhibition in kinase activity of these CDKs and cyclins
could in part be due to an up-regulated level of CDKIs and their
binding with CDK-cyclin complex, causing a resultant G
1
arrest. To
support this anticipation, we immunoprecipitated Cip1/p21 and
Kip1/p21 from total cell lysates, and studied their binding with
CDK2, CDK4, and CDK6, which showed an increase in the bound
levels of these proteins following decursin treatment (Fig. 5D).
These results suggest a possible increase in CDK-cyclin-CDKI
complex formation accompanied by a decrease in CDK kinase
activity following decursin treatment.
Decursin Induces Apoptotic Death of DU145 Cells. During
cell growth assay, we observed that decursin causes a significant
increase in DU145 cell death, where higher doses and longer
treatment times were more effective. Here, using annexin
V-staining and flow cytometry, we assessed whether decursin-
caused cell death is mediated by apoptosis. Cells were treated with
DMSO control, 50, and 100 Amol/L doses of decursin for 24 and 48
hours for the apoptosis analysis. Our data show a dose-dependent
increase (up to f3-fold, P< 0.05-0.001) in apoptotic cell population
following decursin treatment (Fig. 6A). Lower doses (50 Amol/L) of
Figure 3. Effect of decursin on cell cycle progression in DU145 cells. Cells were cultured in complete medium, and treated with either DMSO vehicle control or 25
to 100 Amol/L doses of decursin. After 24 hours of these treatments, cells were collected, washed with PBS, digested with RNase, and then cellular DNA was
stained with propidium iodide as detailed in Materials and Methods. Flow cytometric analysis was then performed for cell cycle distribution. A, propidium iodide
fluorescence pattern for cell cycle distribution in different treatments. B, the percentage of cell cycle distribution data for each treatment group. Mean FSE of three
independent samples. DN, decursin; *, P < 0.05;
$
,P< 0.01;
#
,P< 0.001 versus control.
Cancer Research
Cancer Res 2005; 65: (3). February 1, 2005 1040 www.aacrjournals.org
Research.
on June 3, 2013. © 2005 American Association for Cancercancerres.aacrjournals.org Downloaded from
decursin were not as effective as higher doses (100 Amol/L) which
showed more apoptotic cell death (23%) at 48 hours as compared
with 24 hours of treatment (10%) versus controls showing 5% to 8%
apoptotic cell death (Fig. 6A). These data suggest that apoptosis
induction could be a major mechanism of decursin-caused death of
prostate cancer cells. With similar decursinol treatment, we did not
observe any considerable apoptosis induction in DU145 cells (data
not shown).
Role of Caspase Activation in Decursin-Caused Apoptosis.
Activation of caspase cascade leading to PARP cleavage is
regarded as a major pathway in apoptosis induction (26). Based
on the above results showing induction of apoptosis by decursin,
we analyzed the levels of cleaved caspase-9 and caspase-3
following 48 hours of treatment. Cleavage of caspases is directly
related to their activation status. Our data show that decursin
caused an increase in cleaved caspase-9 and caspase-3, which
were very prominent at 100 Amol/L dose of decursin (Fig. 6B,
rows 1 and 2). Consistent with the cleavage of caspases, decursin
also caused a strong increase in PARP cleavage (Fig. 6B). First,
we used PARP antibody (BD PharMingen) which recognizes both
total (116 kDa) as well as cleaved PARP (89 kDa; Fig. 6B,row 3);
and later the same membrane was stripped and reprobed with
specific cleaved PARP antibody (Cell Signaling) which showed
better sensitivity for the detection of cleaved PARP fragment
(Fig. 6B,row 4). Furthermore, these membranes were also
checked for the level of h-actin as a loading control; only a
representative blot is shown in Fig. 6B(row 5).
To further establish the role of caspase activation in decursin-
caused apoptosis, we used all-caspases inhibitor z-VAD-fmk
(2 hours pretreatment), which only partially reversed the
decursin-induced apoptosis (Fig. 6C). To confirm whether the
dose of all-caspases inhibitor used in the study was sufficient to
inhibit caspase activity, caspase-3 activity was also measured under
identical treatments. As shown in Fig. 6D, 100 Amol/L dose of
caspase inhibitor completely inhibited 100 Amol/L decursin-
induced (f3 fold) caspase-3 activity. Overall, these results
suggested the involvement of both caspase-dependent and
caspase-independent pathway(s) in decursin-induced apoptotic
death of DU145 cells.
Discussion
The central and novel finding in the present study is the
identification of in vitro anticancer efficacy of decursin against
advanced human prostate carcinoma DU145, PC-3, and LNCaP
cells. The completed studies clearly and convincingly show that
decursin causes a G
1
cell cycle arrest via an induction of Cip1/p21
and to a lesser extent Kip1/p27 together with an inhibition in CDK-
cyclin kinase activity as an underlying mechanism in its DU145 cell
growth inhibition. Furthermore, this agent causes apoptosis
induction involving both caspase-dependent and caspase-indepen-
dent mechanisms in DU145 cells. More importantly, decursin was
nontoxic to nonneoplastic human prostate epithelial cells and
showed only a moderate cell growth inhibition with a slight effect
on cell cycle progression.
The present study also provides an evidence for the structure
and anticancer activity of the coumarin compounds decursin
and decursinol against hormone refractory human prostate
cancer DU145 cells. Based on structural analysis, the side chain
in decursin, which is substituted with an -OH group in
decursinol (Fig. 1), could be attributed to the enhanced
anticancer efficacy of decursin. We also observed that decursi-
nol-caused death of DU145 cells was not associated with
apoptosis induction (data not shown); therefore, the substituted
side chain in decursin might be solely responsible for its
apoptotic efficacy in DU145 cells. Overall, the differential efficacy
of these compounds in DU145 cells was in accord with their
structure-activity relationship in inhibiting acetylcholinesterase
activity reported earlier (18).
CDKs, CDKIs, and cyclins play essential roles in the regulation of
cell cycle progression (25). CDKIs are tumor suppressor proteins
that down-regulate the cell cycle progression by binding with active
CDK-cyclin complexes and thereby inhibiting their kinase activities
Figure 4. Effect of decursin on cell cycle progression in PC-3, LNCaP, and
PWR-1E cells. Cells were cultured in complete medium, and treated with either
DMSO vehicle control or 25 to 100 Amol/L doses of decursin. After 24 hours of
these treatments, cells were processed and analyzed by flow cytometry as
mentioned in Fig. 3. The percentage of cell cycle distribution data for PC-3 (A),
LNCaP (B), and PWR-1E (C) cells are shown as mean FSE of three
independent samples. DN, decursin; *, P < 0.05;
$
,P< 0.01;
#
,P< 0.001 versus
control.
A Novel Anticancer Agent Decursin against Prostate Carcinoma Cells
www.aacrjournals.org 1041 Cancer Res 2005; 65: (3). February 1, 2005
Research.
on June 3, 2013. © 2005 American Association for Cancercancerres.aacrjournals.org Downloaded from
(25, 27, 28). Cip1/p21 is a universal inhibitor of CDKs whose
expression is mainly regulated by the p53 tumor suppressor protein
(29); however, Kip1/p27 is up-regulated in response to antiprolifer-
ative signals (30, 31). The increased expression of G
1
cyclins in
cancer cells provides them an uncontrolled growth advantage
because most of these cells either lack CDKI, harbor nonfunctional
CDKI, or CDKI expression is not at a sufficient level to control
CDK-cyclin activity (28, 32). Consistent with these reports, cell cycle
analysis data showed that decursin caused a strong G
1
arrest in cell
cycle progression of prostate cancer cells. Furthermore, mechanis-
tic investigation showed that decursin-induced G
1
arrest in DU145
cells is mainly mediated via an up-regulation of Cip1/p21 (strong)
and Kip1/p27 (moderate). As CDK activity is essential for driving
the cells through G
1
-S transition (33), concomitant with CDKIs
induction, we also observed an increase in Cip1/p21-CDK2/CDK4/
CDK6 and Kip1/p27-CDK2/CDK4/CDK6 binding together with an
inhibition in CDK-cyclin kinase activity in decursin-treated cells.
The increased expression of CDKIs by decursin might also have a
direct relevance in prostate cancer growth and progression, as
decreased Kip1/p27 expression in prostatic carcinomas has been
associated with aggressive phenotype and poor prognosis, and a
failure of irradiation response in prostate cancer patients has been
linked to the loss of Cip1/p21 function (34). Recent studies have
also indicated the prognostic significance of CDKIs in prostate
cancer (35). Specifically, Kip1/p27 expression has been shown to be
an independent predictor of prostate-specific antigen failure
following radical prostatectomy (36), and its low expression is
correlated with poor disease-free survival in prostate cancer
patients (37).
Most of the presently available cytotoxic anticancer drugs
mediate their effect via apoptosis induction in cancer cells
(26, 38), and apoptosis is suggested as one of the major mech-
anisms for the targeted therapy of various cancers including
prostate cancer (26, 38–40). In case of advanced prostate cancer,
cancer cells become resistant to apoptosis and do not respond to
cytotoxic chemotherapeutic agents (7). Therefore, the agents that
induce apoptotic death of hormone-refractory prostate cancer cells
could be useful in controlling this malignancy (41). Consistent with
this approach, our data showing an induction of apoptotic death of
advanced prostate cancer cells by decursin could be of greater
significance in identifying another anticancer mechanism (together
with cell cycle arrest) of decursin for its possible application in
prostate cancer control. Because induction of CDKIs has been
reported in anticancer agent–caused apoptosis in human prostate
cancer cells (42), the decursin-induced CDKIs could be, in part,
responsible for the observed apoptotic death of DU145 cells. The
increase in the levels of active caspase-9 and caspase-3 as well as
cleaved PARP suggested that caspase activation is another
Figure 5. Effect of decursin on G
1
cell cycle regulators and CDK-associated and cyclin-associated kinase activities in DU145 cells. Cells were cultured in complete
medium, and treated with either DMSO vehicle control or 25 to 100 Amol/L doses of decursin as described in Materials and Methods. Aand B, at the end of treatments,
total cell lysates were prepared and subjected to SDS-PAGE followed by Western immunoblotting. Membranes were probed with anti-Cip1/p21, Kip1/p27, CDK2,
CDK4, CDK6, cyclin D1, cyclin E, and h-actin antibodies followed by peroxidase-conjugated appropriate secondary antibodies, and visualized by enhanced
chemiluminescence detection system. The experiments were repeated thrice with similar results and a representative blot is shown for each protein. C, CDK2
and cyclin E kinase activities were determined by in-bead histone H1 kinase assay using immunoprecipitated CDK2 and cyclin E from total cell lysates (200 Ag protein)
using specific antibodies. CDK4, CDK6, and cyclin D1-associated kinase activities were determined by in-bead RB-GST fusion protein kinase assay using
immunoprecipitated CDK4, CDK6, and cyclin D1 from total cell lysates (200 Ag protein) using specific antibodies as described in Materials and Methods. After the assay,
the labeled substrates were subjected to SDS-PAGE, and the gel was dried and exposed to X-ray film. D, Cip1/p21 and Kip1/p27 were immunoprecipitated using
primary specific antibodies, followed by Western blot analysis for CDK2, CDK4, and CDK6 as detailed in Materials and Methods. IP, immunoprecipitation; IB,
immunoblotting.
Cancer Research
Cancer Res 2005; 65: (3). February 1, 2005 1042 www.aacrjournals.org
Research.
on June 3, 2013. © 2005 American Association for Cancercancerres.aacrjournals.org Downloaded from
important mechanism in decursin-induced apoptosis in prostate
cancer cells. However, use of all-caspases inhibitor also showed the
involvement of caspase-independent mechanism(s) in decursin-
mediated apoptosis of DU145 cells. Further studies are needed to
explore the caspase-independent mechanism(s) of apoptosis
induction by decursin.
Several studies suggest that hormone refractory, advanced
prostate cancer phenotype is associated with many changes
including deregulated cell cycle and cell survival signaling, and
inactivation of p53 and pRb such as in DU145 cells. Therefore,
findings in the present study have clinical significance in the fact
that decursin caused p53-independent up-regulation of Cip1/p21
as DU145 cells do not have functional p53 gene and are also
mutated for Rb. Another significant observation was that
decursin showed a sustained G
1
arrest in LNCaP cells, which
have functional p53 and Rb genes, although this effect was
diminished with the increase in treatment time in DU145 and
PC-3 cells lacking functional p53 and pRb. This observation
suggests the most likely role of p53 and/or pRb in decursin-
induced G
1
arrest in LNCaP cells; however, more studies are
needed to confirm this anticipation. Furthermore, we also
observed that decursin induces S and/or G
2
-M arrest at the
expense of G
1
phase cell population with the increase in
treatment time or at 100 Amol/L dose only in PC-3 cells and not
in DU145 or LNCaP cells, suggesting differential molecular
determinants of this cell cycle effect of decursin in PC-3 cells.
In conclusion, our present findings showing the in vitro
anticancer efficacy of decursin, with mechanistic rationale (cell
cycle arrest and apoptosis induction), against advanced human
prostate cancer cells without any cytotoxicity to nonneoplastic
prostate epithelial cells, warrant its further in vivo efficacy studies
in preclinical human prostate cancer models as well as estimation
of pharmacologically achievable doses having biological signifi-
cance in in vitro studies. The positive outcomes of such an in vivo
study could form a strong basis for the development of decursin
as a novel agent for human prostate cancer prevention and/or
intervention.
Acknowledgments
Received 6/4/2004; revised 11/17/2004; accepted 11/22/2004.
Grant support: Supported in part by grants NCI RO1 CA91883 and CA102514.
The costs of publication of this article were defrayed in part by the payment of page
charges. This article must therefore be hereby marked advertisement in accordance
with 18 U.S.C. Section 1734 solely to indicate this fact.
Figure 6. Apoptotic effect of decursin on DU145 cells. A, cells were cultured in complete medium, and treated with either DMSO vehicle control or 25 to 100
Amol/L doses of decursin as described in Materials and Methods. At the end of treatments, total cells were collected and stained with annexin V/propidium iodide as
mentioned in Materials and Methods followed by flow cytometric analysis. Data are presented as a percentage of annexin V/propidium iodide stained cells for each
treatment. B, after 48 hours of decursin treatments, cell lysates were prepared and SDS-PAGE and Western blot analysis were performed for cleaved caspase-9,
cleaved caspase-3, total and cleaved PARP, and h-actin using specific antibodies as described in Materials and Methods. C, cells were cultured in complete medium,
and treated with either DMSO vehicle control, 100 Amol/L all-caspases inhibitor (z-VAD-fmk), or 100 Amol/L dose of decursin and/or all caspase inhibitor 2 hours prior to
decursin in combination treatment for a total 48 hours. At the end of treatments, cells were analyzed for apoptosis as detailed in Materials and Methods. D, in similar
treatment as in (C), cell lysates were prepared and caspase-3 activity assay was performed as detailed in Materials and Methods. Data presented in (A), (C), and (D)
are mean FSE of three independent samples in each treatment group. CI, 100 Amol/L all caspase inhibitor; DN, 100 Amol/L decursin;
#
,P< 0.05;
$
,P< 0.01; *,
P< 0.001 versus control.
A Novel Anticancer Agent Decursin against Prostate Carcinoma Cells
www.aacrjournals.org 1043 Cancer Res 2005; 65: (3). February 1, 2005
Research.
on June 3, 2013. © 2005 American Association for Cancercancerres.aacrjournals.org Downloaded from
References
1. Jemal A, Thomas A, Murray T, et al. Cancer statistics,
2003. CA Cancer J Clin 2003;53:5–26.
2. Bosland MC, McCormick DL, Melamed J, Walden
PD, Jacquotte AZ, Lumey LH. Chemoprevention
strategy for prostate cancer. Eur J Cancer Chemoprev
2002;11:s18–27.
3. Surh YJ. Cancer chemoprevention with dietary phyto-
chemicals. Nat Rev Cancer 2003;3:768–80.
4. Kelloff GJ, Lieberman R, Steele VE, et al. Chemo-
prevention of prostate cancer: concepts and strategies.
Eur Urol 1999;35:342–50.
5. Feldman BJ, Feldman D. The development of andro-
gen-independent prostate cancer. Nat Rev Cancer 2001;
1:34–45.
6. Koivisto P, Kolmer M, Visakorpi T, Kallioniemi OP.
Androgen receptor gene and hormonal therapy failure
of prostate cancer. Am J Pathol 1998;152:1–9.
7. Pilat MJ, Kamradt JM, Pienta KJ. Hormone resis-
tance in prostate cancer. Cancer Metastasis 1998–99;
17:373–81.
8. Agarwal R. Cell signaling and regulators of cell cycle as
molecular targets for prostate cancer prevention by
dietary agents. Biochem Pharmacol 2000;60:1051–9.
9. Hong WK, Sporn MB. Recent advances in chemo-
prevention of cancer. Science 1997;278:1073–7.
10. Kucuk O. Chemoprevention of prostate cancer.
Cancer Metastasis Rev 2002;21:111–24.
11. Singh RP, Agarwal R. Prostate cancer prevention by
silibinin. Curr Cancer Drug Targets 2004;4:1–11.
12. Giovannucci E. Nutrition, insulin, insulin-like growth
factors and cancer. Horm Metab Res 2003;35:694–704.
13. Barnes S. Role of phytochemicals in prevention and
treatment of prostate cancer. Epidemiol Rev 2001;
23:102–5.
14. Singh RP, Dhanalakshmi S, Agarwal R. Phytochem-
icals as cell cycle modulators: a less toxic approach in
halting human cancer. Cell Cycle 2002;1:156–61.
15. Nelson PS, Montgomary B. Unconventional therapy
for prostate cancer: good, bad or questionable? Nat Rev
Cancer 2003;3:845–58.
16. Denmeade SR, Isaacs JT. A history of prostate cancer
treatment. Nat Rev Cancer 2002;2:389–96.
17. Chi HJ, Kim HS. Studies on the components of
Umbelliferae plants in Korea: pharmacological study
of decursin, decursinol and nodakenin. Korean
J Pharmacog 1970;1:25–32.
18. Kang SY, Lee KY, Sung SH, Park MJ, Kim YC.
Coumarins isolated from Angelica gigas inhibit acetyl-
cholinesterase: structure-activity relationship. J Nat
Prod 2001;64:683–5.
19. Lee S, Shin DS, Kim JS, Oh KB, Kang SS. Antibacterial
coumarins from Angelica gigas roots. Arch Pharm Res
2003;26:449–52.
20. Ahn KS, Sim WS, Kim IH. Decursin: a cytotoxic agent
and protein kinase C activator from the root of Angelica
gigas. Planta Med 1996;62:7–9.
21. Lee S, Lee YS, Jung SH, Shin KH, Kim BK, Kang SS.
Antitumor activity of decursinol angelate and decursin
from Angelica gigas. Arch Pharm Res 2003;26:727–30.
22. Lee S, Kang SS, Shin KS. Coumarins and pyrimidine
from Angelica gigas roots. Nat Product Sci 2002;8:58–61.
23. Singh RP, Agarwal C, Agarwal R. Inositol hexa-
phosphate inhibits growth and induces G
1
arrest and
apoptotic death of human prostate carcinoma DU145
cells: modulation of CDKI-CDK-cyclin and pRb-
related protein-E2F complexes. Carcinogenesis 2003;
24:555–63.
24. Agarwal C, Singh R, Agarwal R. Grape seed extract
induces apoptotic death of human prostate carcino-
ma DU145 cells via caspases activation accompanied
by dissipation of mitochondrial membrane potential
and cytochrome crelease. Carcinogenesis 2002;23:
1869–76.
25. Grana X, Reddy P. Cell cycle control in mammalian
cells: role of cyclins, cyclin-dependent kinases (CDKs),
growth suppressor genes and cyclin-dependent kinase
inhibitors (CDKIs). Oncogene 1995;11:211–9.
26. Lowe SW, Lin AW. Apoptosis in cancer. Carcino-
genesis 2000;21:485–95.
27. Morgan DO. Principles of CDK regulation. Nature
(Lond) 1995;374:131–4.
28. Hunter T, Pine J. Cyclins and cancer II: cyclin D and
CDK inhibitors come of age. Cell 1994;79:573–82.
29. Xiong Y, Hannon GJ, Zhang H, Casso D, Kobayashi R,
Beach D. P21 is a universal inhibitor of cyclin kinases.
Nature 1993;366:701–4.
30. Toyoshima H, Hunter T. P27, a novel inhibitor of G
1
cyclin-CDK protein kinase activity, is related to p21. Cell
1994;78:67–74.
31. Polyak K, Lee MH, Erdjument-Bromage H, et al.
Cloning of Kip1/p27, cyclin-dependent kinase inhibitor
and a potential mediator of extracellular antimitogenic
signals. Cell 1994;78:59–66.
32. Dulic V, Lees E, Reed SI. Association of human cyclin
E with a periodic G
1
-S phase protein kinase. Science
(Washington DC) 1992;257:1958–61.
33. Tsai LH, Lees E, Faha B, Harlow E, Riabowol K. The
CDK2 kinase is required for the G
1
to S transition in
mammalian cells. Oncogene 1993;8:1593–602.
34. Cheng L, Lloyd RV, Weaver AL, et al. The cell cycle
inhibitors p21/waf1 and Kip1/p27 are associated with
survival in patients treated by salvage prostatectomy
after radiation therapy. Clin Cancer Res 2000;6:1896–9.
35. Tsihlias J, Kapusta L, Slingerland J. The prognostic
significance of altered cyclin-dependent kinase inhib-
itors in human cancers. Ann Rev Med 1999;50:401–23.
36. Yang RM, Naitoh J, Murphy M, et al. Low p27
expression predicts poor disease-free survival in
patients with prostate cancer. J Urol 1998;159:941–5.
37. Freedland SJ, deGregorio F, Sacoolodge JC, et al.
Preoperative p27 status is an independent predictor of
prostate specific antigen failure following radical
prostatectomy. J Urol 2003;169:1325–30.
38. Gurumurthy S, Vasudevan KM, Rangnekar VM.
Regulation of apoptosis in prostate cancer. Cancer
Metastasis Rev 2001;20:225–43.
39. Guseva NV, Taghiyev AF, Rokhlin OW, Cohen MB.
Death receptor-induced cell death in prostate cancer.
J Cell Biochem 2004;91:70–99.
40. Jiang C, Wang Z, Ganther H, Lu J. Caspases as key
executors of methyl selenium-induced apoptosis
(anoikis) of DU145 prostate cancer cells. Cancer Res
2001;61:3062–70.
41. Kantoff PW. New agents in the therapy of hormone-
refractory prostate cancer. Semin Oncol 1995;22:32–4.
42. Don MJ, Chang YH, Chen KK, Ho LK, Chau YP.
Induction of CDK inhibitors (p21/waf1 and Kip1/p27)
and Bak in the h-lapachone-induced apoptosis of
human prostate cancer cells. Mol Pharmacol 2001;
59:784–94.
Cancer Research
Cancer Res 2005; 65: (3). February 1, 2005 1044 www.aacrjournals.org
Research.
on June 3, 2013. © 2005 American Association for Cancercancerres.aacrjournals.org Downloaded from
... Yim et al. [56], conducted a more thorough analysis of compound (11) and its effects on the androgen receptor-positive LNCaP cell line as well as on the androgen receptor-negative prostate cancer cell lines (DU145 and PC-3). Their results showed that this compound, at concentrations between 25 and 100 μM, effectively inhibited the development of these cell lines. ...
... aversive memory performance produced by beta-amyloid peptide 1-42 in mice preconditioned for four weeks at 0.001%, 0.002%, and 0.004% concentrations, according to Yan et al. [140]. Furthermore, without influencing total locomotor attribute, compound (56), at a dosage of 0.004%, effectively countered the decline in alternation behavior (connected to navigational memory) caused by beta-amyloid. These data imply that compound (56) when taken as a preventive strategy, may be able to shield AD patients' memory from the deficits caused by beta-amyloid. ...
Article
Heterocycle conjugates provide a fresh investigative scope to find novel molecules with enhanced phytothera-peutic characteristics. Coumarin-based products are widely used in the synthesis of several compounds with biological and medicinal properties since they are naturally occurring heterocycles with a broad dispersion. The investigation of coumarin-based phytochemicals with annulated heterocyclic rings is a promising approach to discovering novel conjugates with significant phytotherapeutic attributes. Due to the applicable coumarin extraction processes, a range of linear coumarin-heterocyclic conjugates were isolated from different natural resources and exhibited remarkable therapeutic efficacy. This review highlights the phytotherapeutic potential and origins of various natural linear coumarin-heterocyclic conjugates. We searched several databases, including Science Direct , Web of Science, Springer, Google Scholar, and PubMed. After sieving, we ultimately identified and included 118 pertinent studies published between 2000 and the middle of 2023. This will inspire medicinal chemists with extremely insightful ideas for designing and synthesizing therapeutically active lead compounds in the future that are built on the pharmacophores of coumarin-heterocyclic conjugates and have significant therapeutic attributes .
... Decursinol angelate was able to significantly inhibit the invasion of the fibrosarcoma cell line HT1080 and the breast cancer cell line MDA-MB-231, as well as inhibit the increase of B16F10 melanoma cells (Kim et al., 2010c;Chang et al., 2021). Second, decursinol is not only the active derivative of decursin but also its metabolite (Yim et al., 2005;Lee et al., 2009). The side chain of decursin (CH3)2-C = CH-COO is replaced by -OH, but decursinol, which has the same coumarin ring system, is not as anticancer as decursin, suggesting that this isoprenoid structure may be important for the anticancer effect (Kim et al., 2005;Yim et al., 2005;Song et al., 2007). ...
... Second, decursinol is not only the active derivative of decursin but also its metabolite (Yim et al., 2005;Lee et al., 2009). The side chain of decursin (CH3)2-C = CH-COO is replaced by -OH, but decursinol, which has the same coumarin ring system, is not as anticancer as decursin, suggesting that this isoprenoid structure may be important for the anticancer effect (Kim et al., 2005;Yim et al., 2005;Song et al., 2007). Interestingly, the Son team further showed that oral administration of decursinol at 10 mg/kg resulted in a significantly better reduction of lung tumor nodules in mice than decursin at the same dose (Son et al., 2011). ...
Article
Full-text available
Cancer is a globally complex disease with a plethora of genetic, physiological, metabolic, and environmental variations. With the increasing resistance to current anticancer drugs, efforts have been made to develop effective cancer treatments. Currently, natural products are considered promising cancer therapeutic agents due to their potent anticancer activity and low intrinsic toxicity. Decursin, a coumarin analog mainly derived from the roots of the medicinal plant Angelica sinensis, has a wide range of biological activities, including anti-inflammatory, antioxidant, neuroprotective, and especially anticancer activities. Existing studies indicate that decursin affects cell proliferation, apoptosis, autophagy, angiogenesis, and metastasis. It also indirectly affects the immune microenvironment and can act as a potential anticancer agent. Decursin can exert synergistic antitumor effects when used in combination with a number of common clinical anticancer drugs, enhancing chemotherapy sensitivity and reversing drug resistance in cancer cells, suggesting that decursin is a good drug combination. Second, decursin is also a promising lead compound, and compounds modifying its structure and formulation form also have good anticancer effects. In addition, decursin is not only a key ingredient in several natural herbs and dietary supplements but is also available through a biosynthetic pathway, with anticancer properties and a high degree of safety in cells, animals, and humans. Thus, it is evident that decursin is a promising natural compound, and its great potential for cancer prevention and treatment needs to be studied and explored in greater depth to support its move from the laboratory to the clinic.
... Cumarin includes pyranocoumarins (decursin and decursinol algelate), umbelliferone, nodakenin, peucedanone, marmesin, demethylsuberosin and isoimperatorin 23) . Decursin and decursinol algelate are the main active components of the biological effects of AGN 23,24) . AGN root contains more than 6% of the active components, decursin, decursinol angelate and nodakenin, in particular, decursin and decursinol angelate are not found in Chinese and Japanese Angelica species 25) . ...
... In general, the cell cycle arrest is a prestage that induces apoptosis. Previous reports have reported that decursin acts on the cell cycle and is involved in antitumor efects in prostate cancer, bladder cancer, and colorectal cancer [29,33]. In this study, decursin increased the number of cells in the G0/G1 phase and decreased in the S-phase on human osteosarcoma cell lines at 24 h. ...
Article
Full-text available
Osteosarcoma is a rare malignant tumor that commonly occurs in children. Anticancer drugs, for example, cisplatin, aid in postsurgery recovery but induce side effects such as renal damage, affecting the life prognosis of patients. Decursin which is one of the bioactive components has been reported for its anti-inflammatory, antioxidant, and antitumor effects, but the effect on osteosarcoma is unexplained. In this study, the research theme was to examine the sensitizing effect of decursin and its influence on cisplatin-induced nephrotoxicity. The cell viability and half maximal inhibitory concentration (IC50), apoptosis induction, and effect on cell cycle and Akt pathways were examined. In vivo, we examine the effects of decursin on tumors and mice bodies. Additionally, the effects of the cisplatin-decursin combination were evaluated in vitro and in vivo. Decursin suppressed cell viability and induced apoptosis via the cell cycle. Decursin also inhibited the Akt pathway by suppressing the phosphorylation of Akt. It enhanced apoptosis induction and lowered cell viability in combination with cisplatin. The increasing tumor volume was suppressed in the decursin-administrated group with further suppression in combination with cisplatin compared to sole cisplatin administration. The decrease in renal function and renal epithelial cell damage caused by cisplatin was improved by the combinatorial treatment with decursin. Therefore, decursin demonstrated an antitumor effect on the osteosarcoma cells and a renal protective effect in combination with cisplatin. Therefore, decursin is a prospective therapeutic agent against osteosarcoma.
... Essential oils exhibit a very interesting chemotherapeutic potential; several essential oil constituents have been described as cytotoxic agents comprising β-caryophyllene, β-elemene, δ-elemene, α-humulene and others (Wang et al., 2005;Sylvestre et al., 2006;Hou et al., 2006;Tao et al., 2006;Xiao et al., 2006). Concerning essential oilbearing species of the Apiaceae, many of these have been shown to exert unique cytotoxic and antileukaemic activities (Babu et al., 1995;Pae et al., 2002;Yim et al., 2005). The genus Pituranthos ( Apiaceae) is represented in Egypt by two species of wild growing desert shrubs, P. tortuosus (Desf.) ...
Article
Full-text available
The aerial parts of Pituranthos tortuosus (Desf.) Benth and Hook (Apiaceae), growing wild in Egypt, yielded 0.8%, 0.6%, and 1.5% (v/w) of essential oil when prepared by hydrodistillation (HD), simultaneous hydrodistillation-solvent (n-pentane) extraction (Lickens- Nickerson, DE), and conventional volatile solvent extraction (preparation of the “absolute”, SE), respectively. GC-MS analysis showed that the major components in the HD sample were β-myrcene (18.81%), sabinene (18.49%), trans-iso-elemicin (12.90%), and terpinen- 4-ol (8.09%); those predominent in the DE sample were terpinen-4-ol (29.65%), sabinene (7.38%), γ-terpinene (7.27%), and β-myrcene (5.53%); while the prominent ones in the SE sample were terpinen-4-ol (15.40%), dill apiol (7.90%), and allo-ocimene (4E,6Z) (6.00%). The oil prepared in each case was tested for its cytotoxic activity on three human cancer cell lines, i.e. liver cancer cell line (HEPG2), colon cancer cell line (HCT116), and breast cancer cell line (MCF7). The DE sample showed the most potent activity against the three human cancer cell lines (with IC50 values of 1.67, 1.34, and 3.38 μg/ml against the liver, colon, and breast cancer cell lines, respectively). Terpinen-4-ol, sabinene, γ-terpinene, and β-myrcene were isolated from the DE sample and subjected to a similar evaluation of cytotoxic potency; signifi cant activity was observed
Article
Full-text available
Cancers are currently the major cause of mortality in the world. According to previous studies, matrix metalloproteinases (MMPs) have an impact on tumor cell proliferation, which could lead to the onset and progression of cancers. Therefore, regulating the expression and activity of MMPs, especially MMP-2 and MMP-9, could be a promising strategy to reduce the risk of cancers. Various studies have tried to investigate and understand the pathophysiology of cancers to suggest potent treatments. In this review, we summarize how natural products from marine organisms and plants, as regulators of MMP-2 and MMP-9 expression and enzymatic activity, can operate as potent anticancer agents.
Article
The hybridization of heterocycles presents a key opportunity to craft innovative multicyclic compounds with enhanced biological activity. Coumarins, being broadly prevalent natural heterocycles, are extensively utilized in the formulation of diverse biologically and pharmacologically active chemicals. The fusion of various hetero rings with the coumarin ring represents a captivating approach to creating novel hybrid molecules endowed with notable biological activities. In the endeavor of developing heterocyclic-fused coumarins, a diverse array of 6,7-heterocycle-coumarin hybrids has been introduced, showcasing remarkable biological efficacy. The impact of heterocyclic annulation at the 6,7-positions of the coumarin ring on the biological activity of the resultant structures has been examined. This review centers on the natural origins, synthetic methodologies, structural activity relationship investigation, and biological potentials of 6,7-heterocycle-coumarin hybrids. We conducted searches across several databases, including Web of Science, Google Scholar, PubMed, and Scopus. After sieving, we ultimately identified and included 161 pertinent studies published between 1995 and the middle of 2023. This will offer valuable insight to medicinal chemists for the future design and synthesis of biologically active lead compounds based on heterocycle-fused coumarin scaffolds with substantial therapeutic effects.
Article
Angelica species have been traditionally used for their medicinal properties. Recent studies have suggested their potential use as anticancer agents, making them an area of interest for further research. The review aims to summarize the current understanding of the potential anticancer effects of Angelica species and to provide insights for further research in this area. We searched for "Angelica" related information on Google Scholar, PubMed, ScienceDirect, Wiley, Science Citation Index Finder, and Springer link by searching keywords such as "Angelica," "Angelica phytochemical," "Angelica antitumor effect," "Angelica molecular mechanisms," and "Angelica clinical application." Included articles focused on the Angelica plant's anticancer properties and clinical studies, while non-cancer-related biological or phytochemical investigations were excluded. We conducted a comprehensive search of books, journals, and databases published between 2001 and 2023, identifying 186 articles for this narrative review. The articles were analyzed for their potential anticancer properties and therapeutic applications. Active compounds in the Angelica genus, such as coumarins, furanocoumarins, phthalides, and polysaccharides, exhibit anticancer properties through various mechanisms. Specific species, like A. archangelica, Angelica sinensis, A. gigas, and A. ksiekie, have the potential as anticancer agents by targeting cellular pathways, generating reactive oxygen species, and inducing apoptotic cell death. Further research into the properties of the Angelica genus is needed for developing new treatments for cancer. Phytochemicals from Angelica species possess potential as anticancer agents, requiring further research for the development of effective, low-cost, and low-toxicity cancer treatments compared to synthetic antitumor drugs.
Article
We have shown that decursin, a coumarin compound, induces cell cycle arrest and apoptosis in human prostate cancer cells (PCa); however, its molecular mechanisms are largely unexplored. We studied the mechanisms associated with its anticancer activity in advanced human prostate carcinoma cells. We found that decursin inhibited epidermal growth factor receptor (EGFR) signaling by inhibiting its activating phosphorylation at tyrosine 1068 residue in DU145 and 22Rv1 cells. This inhibition of EGFR was associated with the downregulation of ERK1/2 phosphorylation. Both EGFR and ERK1/2 are known to be deregulated/activated in many human malignancies. Consistent with our earlier study, decursin (25–100 µM) treatment for 24–72 h inhibited DU145 cell proliferation by 49%–87% (p < 0.001) which was associated with strong G1 phase arrest and cell death. It also decreased (p < 0.001) the number of surviving colonies. Decursin moderately increased the expression of Rb‐related proteins p107 and p130 but decreased the levels of E2F family transcription factors including E2F‐3, E2F‐4 and E2F‐5. Further, decursin strongly inhibited the growth of androgen–dependent prostate carcinoma 22Rv1 cells from 61% to 79% (p < 0.001) and arrested these cells at G1 phase via induction of cyclin‐dependent kinase inhibitor p27/Kip1 and downregulation of CDK2 and CDK4 protein expression. Additionally, EGFR inhibitor erlotinib‐ and EGF ligand‐modulated EGFR activation validated EGFR signaling as a target of decursin‐mediated cell growth inhibition and cytotoxicity. Decursin decreased EGF ligand‐induced phosphorylation of EGFR (Y‐1068) as well as activation of its downstream mediator, ERK1/2. Furthermore, inhibitory targeting of EGFR‐ERK1/2 axis by combinatorial treatment of decursin and erlotinib further sensitized DU145 cells for the decursin‐induced growth inhibition and cell death. Overall, these findings strongly suggest that anticancer efficacy of decursin against human PCa involves inhibitory targeting of EGFR‐ERK1/2 signaling axis, a pathway constitutively active in advanced PCa.
Article
Introduction Medicinal plants and their highly valuable compounds are used worldwide as therapeutic interventions for the management of complex, multifactorial physiologic imbalances, and various other health problems including cancer. However, the characteristics of the active components and the precise mechanisms underlying beneficial effects of medicinal plants are largely unclear, which obstructs their progression to clinical trials and ultimately pharmacy. This integrative review presents critical and synthesized literature on the characteristics and mechanisms of anticancer plants and their derivatives. Method Databases such as PubMed and Google Dataset were searched using key words related to the fields of botanical derivatives and cancer to identify articles that describe the anticancer properties and signal transduction pathways of medicinal plants and their derivatives. Furthermore, the Clinical Trial database was searched to find anticancer drugs that are derived from medicinal plants or natural sources. Publications using this search were gathered based on their relevance to anticancer botanical derivatives and summarized based on their results and significance. Results Anticancer ingredients such as phenolic compounds, flavonoids and alkaloids are present in medicinal plants. These bioactive compounds modulate proliferative and apoptotic factors to prevent tumor growth. Botanical derivatives have been utilized as lead compounds for the development of new anticancer drugs. Several chemotherapy medicines derived from botanical leads have been approved by the FDA. The characteristics that impacted signal transduction pathways and pharmacological application of botanical derivatives are integrated and summarized. Conclusion There are a variety of valuable plant-derived bioactive compounds with anticancer property. Signal cascades that are responsible for proliferation, cell cycle arrest, apoptosis, and angiogenesis are implicated in the actions and effects of medicinal plants and their derivatives. It is expected that novel therapeutics derived from medicinal plants can be developed to treat patients with cancer.
Article
Full-text available
Prostate cancer mortality results from metastasis and is often coupled with progression from androgen-dependent to androgen-independent growth. Unfortunately, no effective treatment for metastatic prostate cancer increasing patient survival is available. The absence of effective therapies reflects in part a lack of knowledge about the molecular mechanisms involved in the development and progression of this disease. Apoptosis, or programmed cell death, is a cell suicide mechanism that enables multicellular organisms to regulate cell number in tissues. Inhibition of apoptosis appears to be a critical pathophysiological factor contributing to the development and progression of prostate cancer. Understanding the mechanism(s) of apoptosis inhibition may be the basis for developing more effective therapeutic approaches. Our understanding of apoptosis in prostate cancer is relatively limited when compared to other malignancies, in particular, hematopoietic tumors. Thus, a clear need for a better understanding of apoptosis in this malignancy remains. In this review we have focused on what is known about apoptosis in prostate cancer and, more specifically, the receptor/ligand-mediated pathways of apoptosis as potential therapeutic targets.
Article
Full-text available
Peer Reviewed http://deepblue.lib.umich.edu/bitstream/2027.42/44511/1/10555_2004_Article_203064.pdf
Article
Full-text available
G1 cyclins control the G1 to S phase transition in the budding yeast, Saccharomyces cerevisiae. Cyclin E was discovered in the course of a screen for human complementary DNAs that rescue a deficiency of G1 cyclin function in budding yeast. The amounts of both the cyclin E protein and an associated protein kinase activity fluctuated periodically through the human cell cycle; both were maximal in late G1 and early S phases. Cyclin E-associated kinase activity was correlated with the appearance of complexes containing cyclin E and the cyclin-dependent kinase Cdk2. Thus, the cyclin E-Cdk2 complex may constitute a human G1-S phase-specific regulatory protein kinase.
Article
Five coumarins and a pyrimidine were isolated from the roots of Angelica gigas. Their structures were elucidated as bergapten (1), decursinol angelate (2), decursin (3), nodakenetin (4), uracil (5) and nodakenin (6) by spectral analysis. Among them, bergapten (1) and uracil (5) were isolated for the first time from this plant part.
Article
Cancer chemopreventive effects of inositol hexaphosphate (IP6), a dietary constituent, have been demonstrated against a variety of experimental tumors, however, limited studies have been done against prostate cancer (PCA), and molecular mechanisms are not well defined. In the present study, we investigated the growth inhibitory effect and associated mechanisms of IP6 in advanced human PCA cells. Advanced human prostate carcinoma DU145 cells were used to study the anticancer effect of IP6. Flow cytometric analysis was performed for cell cycle progression and apoptosis studies. Western immunoblotting, immunoprecipitation and kinase assay were performed to investigate the involvement of G1 cell cycle regulators and their interplay, and end point markers of apoptosis. A significant dose- as well as time-dependent growth inhibition was observed in IP6-treated cells, which was associated with an increase in G1 arrest. IP6 strongly increased the expression of CDKIs (cyclin-dependent kinase inhibitors), Cipl/ p21 and Kipl/p27, without any noticeable changes in G1 CDKs and cyclins, except a slight increase in cyclin D2. IP6 inhibited kinase activities associated with CDK2, 4 and 6, and cyclin E and D1. Further studies showed the increased binding of Kipl/p27 and Cipl/p21 with cyclin D1 and E. In down-stream of CDKI-CDK/cyclin cascade, IP6 increased hypophosphorylated levels of Rb-related proteins, pRb/p107 and pRb2/p130, and moderately decreased E2F4 but increased its binding to both pRb/ p107 and pRb2/p130. At higher doses and longer treatment times, IP6 caused a marked increase in apoptosis, which was accompanied by increased levels of cleaved PARP and active caspase 3. IP6 modulates CDKI-CDK-cyclin complex, and decreases CDK-cyclin kinase activity, possibly leading to hypophosphorylation of Rb-related proteins and an increased sequestration of E2F4. Higher doses of IP6 could induce apoptosis and that might involve caspases activation. These molecular alterations provide an insight into IP6-caused growth inhibition, G1 arrest and apoptotic death of human prostate carcinoma DU145 cells.
Article
Acetylcholinesterase (AChE) inhibitory activity−guided fractionation of Angelica gigas led to isolation and identification of a new coumarin, peucedanone (12), and isolation of 11 known coumarins. Among them, decursinol (1) represented the highest inhibitory activity toward AChE in vitro. The correlation of the inhibitory activities of the coumarins toward AChE with their chemical structures was studied.
Article
Chemoprevention refers to the use of agents to inhibit, reverse or retard tumorigenesis. Numerous phytochemicals derived from edible plants have been reported to interfere with a specific stage of the carcinogenic process. Many mechanisms have been shown to account for the anticarcinogenic actions of dietary constituents, but attention has recently been focused on intracellular-signalling cascades as common molecular targets for various chemopreventive phytochemicals.