Jurnal Nutri3

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Hindawi Publishing Corporation

BioMed Research International


Volume 2013, Article ID 824563, 10 pages
http://dx.doi.org/10.1155/2013/824563

Review Article
Mechanisms of Omega-3 Polyunsaturated Fatty Acids in
Prostate Cancer Prevention

Zhennan Gu,1,2 Janel Suburu,2 Haiqin Chen,1 and Yong Q. Chen1,2


1
State Key Laboratory of Food Science and Technology, School of Food Science and Technology, Jiangnan University,
Wuxi 214122, China
2
Department of Cancer Biology, Wake Forest University School of Medicine, Medical Center Boulevard, Winston-Salem,
NC 27157, USA

Correspondence should be addressed to Yong Q. Chen; [email protected]

Received 6 March 2013; Revised 2 May 2013; Accepted 8 May 2013

Academic Editor: Gabriella Calviello

Copyright © 2013 Zhennan Gu et al. This is an open access article distributed under the Creative Commons Attribution License,
which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

This review focuses on several key areas where progress has been made recently to highlight the role of omega-3 polyunsaturated
fatty acid in prostate cancer prevention.

1. Introduction required enzymes and, therefore, PUFA must be obtained


from the diet. The n-6 and n-3 PUFA also cannot be intercon-
The health benefits of omega-3 polyunsaturated fatty acids verted in mammals, but within each series, their metabolism
(n-3 PUFA), mainly eicosapentaenoic acid (EPA 20:5) and can produce various lipids that differ in chain length and
docosahexaenoic acid (DHA, 22:6), have been long known. number of double bonds. Linoleic acid (LA, 18:2 n-6) is an n-6
Epidemiologic studies dating back to the 1970s were among PUFA found in high concentration in grains as well as many
the first to suggest that dietary PUFA may be beneficial seeds and meats. LA serves as a substrate to be converted
in preventing disease [1, 2]. Still today, studies continue to into a longer fatty acid, arachidonic acid (AA, 20:4 n-6), via
demonstrate the health benefits of n-3 PUFA; however, the a series of oxidative desaturation and elongation reactions.
mechanisms of action of n-3 PUFA are still not fully under- Of the n-3 fatty acids, alpha linolenic acid (ALA, 18:3 n-3)
stood. Many new discoveries have advanced our understand- is found at moderate levels in plants, seeds, leafy vegetables,
ing about the activities of n-3 PUFA against human disease. legumes, and nuts. ALA is not metabolized efficiently to
For example, DHA-receptor GPR120 has been demonstrated longer-chain n-3 PUFA, such as EPA and DHA.
to play a role in sensing and controlling obesity and metabolic Although they belong to two distinct families, n-3 and
syndrome [3]; the recently identified omega-3 mediators, n-6 PUFA are metabolized by some of the same enzymes,
resolvins, and protectins have been demonstrated to have specifically, delta-5-desaturase and delta-6-desaturase. Excess
anti-inflammatory and proresolving activities [4]. The pur- in one family of fatty acid can interfere with the metabolism of
pose of this review is to highlight the recent advances in the other and alter their overall biological effects [5]. During
our understanding of the mechanisms by which n-3 PUFA n-6 PUFA conversion, delta-6-desaturase, or fatty acid desat-
modulate prostate cancer development. urase 2 (FADS2), converts LA to gamma-linolenic acid (GLA,
18:3 n-6). This enzyme represents a rate-limiting step in the
2. Fatty Acids synthesis of AA from LA [6]. GLA is elongated to dihomo-
gamma-linolenic acid (DGLA, 20:3 n-6) through a chain
There are two major classes of PUFA: n-6 and n-3. Unlike reaction of four enzymes: a condensation reaction of the
saturated and monounsaturated fatty acid, PUFA cannot fatty acyl chain with malonyl-CoA, catalyzed by an enzyme
be synthesized de novo by mammals because they lack the encoded by the ELOVL5 gene (elongation of very long-chain
2 BioMed Research International

fatty acids, family member 5); a reduction reaction mediated a Swedish research group found that n-6 PUFA and n-6
by 3-ketoacyl-CoA reductase (KAR); a dehydration reaction PUFA precursors were significantly higher in malignant
catalyzed by 3-hydroxyacyl-CoA dehydratase (HACD); and tissues. This finding further demonstrates that n-6 dietary
a second reduction reaction catalyzed by trans-2,3-enoyl- fat is associated with prostate carcinogenesis [28]. In race-
CoA reductase (TECR). Finally, DGLA is converted to AA specific analyses based on a case-control study comprising
by delta-5 desaturase, or fatty acid desaturase 1 (FADS1) [6, 7]. 79 prostate cancer cases and 187 controls, Williams and
Interestingly, malonyl-CoA, which is necessary for fatty acid colleagues found that a high ratio of n-6 to n-3 fatty acids
elongation, is derived from the rate-limiting enzyme of the may increase the overall risk of prostate cancer among white
de novo fatty acid synthesis pathway, acetyl-CoA carboxylase. men and possibly increase the risk of high-grade prostate
Fatty acid synthesis is well described as an overactive pathway cancer among all men [29].
in many cancers [8–11], and its upregulation may also con- At the same time, epidemiological literature on the
tribute to the elongation of PUFA. association of n-3 PUFA and cancer, including correlational
In contrast to AA, the efficiency of ALA conversion to studies and migrational studies, suggest a protective role
DHA appears to be very low, below 5% in humans. Most played by n-3 PUFA. In a recent population-based prospec-
ingested ALA is subject to beta-oxidation to provide energy, tive cohort study of 90,296 Japanese subjects, Sawada et al.
and only a small fraction is converted to EPA [12, 13]. It reported that consumption of n-3-rich fish or n-3 PUFA,
was estimated that as low as 0.2% of ALA is converted to particularly EPA, DPA, and DHA, appears to protect against
EPA, 64% of EPA to docosapentaenoic acid (DPA, 22:5 n-3), the development of hepatocellular carcinoma (HCC) [32]. In
and 37% of DPA to DHA [14]. Thus, the overall amount another population-based prospective study in Japan, there
of DPA and DHA converted from ALA is about 0.13% was an inverse relationship between marine n-3 PUFA intake
and 0.05% of the starting ALA, respectively. These findings and the risk of colorectal cancer, but this association was
suggest that any contributions from the fatty acid synthesis only statistically significant in the proximal site of the large
pathway toward PUFA metabolism most likely favor n-6, bowel [33]. Chavarro et al. performed a nested case-control
rather than n-3, PUFA elongation. It is also very likely that study by analyzing blood samples of 14,916 healthy men and
synthesis of the longer n-3 fatty acids from ALA within the concluded that higher blood levels of long-chain n-3 fatty
body is competitively hindered by the n-6 analogues. It has acids were associated with a reduced risk of prostate cancer
been reported that the n-3 conversion efficiency is greater in [34]. Szymanski et al. conducted a meta-analysis of fish intake
women, possibly because of the importance of meeting the and prostate cancer by focusing on the incidence of prostate
DHA demands of the fetus and neonate [14]. cancer and prostate cancer-specific mortality. Their results
did not establish a protective association of fish consumption
with prostate cancer incidence but showed a significant 63%
3. PUFA and Cancer reduction in prostate cancer-specific mortality [35].
The results of correlational studies are mixed, some
Total fat intake and the ratio of n-6 to n-3 PUFA in the of them failing to demonstrate a statistically significant
Western diet have increased significantly since the Industrial effect. Several confounding factors could account for the
Revolution [15, 16]. Increased fat consumption has been asso- inconsistent results on the association between n-3 PUFA
ciated with the development of specific types of cancer such and prostate cancer. First, population-based studies mainly
as breast, colon, and pancreatic and prostate cancers, with the rely on data from self-reported dietary fatty acid intake or
notable exception of n-3 PUFA, which show protective effects from estimates based on national consumption, and these
against colon, breast, and prostate cancers in a number of assessments correlate poorly with direct measurements of
experimental systems [17–23]. Epidemiological studies about fatty acids in patient samples. In addition, the actual intake
the association of dietary fat and cancer suggests a protective in n-3 PUFA may be too low for a protective effect in some
effect of n-3 PUFA and a promoting effect of n-6 PUFA on cases. Second, the ratio of n-6 to n-3 fatty acids may be
cancer. Most clinical data regarding the effects of dietary fat more important than the absolute amount of n-3 PUFA,
on cancers are observational [24], and the results of such as suggested by animal and human studies [16, 36]. Using
studies are mixed, as many fail to demonstrate a significant a prostate-specific Pten knockout mouse prostate cancer
association between n-3 PUFA and reduced prostate cancer model, we showed that a ratio of n-6 to n-3 below 5 was
risk or tumor growth [20, 25–27]. effective in slowing cancer progression [3]. Brown et al.
The Western diet contains disproportionally high reported that AA might potentiate the risk of metastatic
amounts of n-6 PUFA and low amounts of n-3 PUFA, prostate cancer cell migration and seeding at the secondary
denoted as a high n-6 to n-3 PUFA ratio. Most data regarding site in vivo, and lowering the n-6/n-3 ratio in diet by uptake
the effects of high dietary n-6 PUFA are positively associated of n-3 PUFA might reduce this risk [37].
with prostate cancer incidence [28–30]. In a study of
Jamaican men undergoing prostate biopsy for elevated PSA
levels, a positive correlation was observed between n-6 fatty 4. Mechanisms of Action
acid LA and Gleason score and n-6 (LA) to n-3 (DHA)
ratio in erythrocyte membranes and prostate tumor volume 4.1. Integration of PUFAs into Plasma Membrane Glycerophos-
[31]. By comparing PUFA content from malignant and pholipids. Although fatty acids are consumed at high levels
benign prostatic tissues from the same prostate specimens, in a typical Western diet, tumor cells display a strong
BioMed Research International 3

dependence on de novo fatty acid synthesis [9, 10]. The opposing the action of PI3K [46, 47]. AKT can also be phos-
increased proliferation and metabolism of cancer cells could phorylated and activated by phosphoinositide-dependent
be the trigger for the abnormal requirement for fatty acid kinase-1 (PDPK1), a PH domain-containing kinase down-
compared to normal cells. Most newly synthesized fatty acids stream of PI3K (for review, see Franke, 2008) [48]. Hence,
are used to support membrane biogenesis in the form of intracellular PIP3 plays a pivotal role in this PI3K/PIP3 /AkT
glycerophospholipids, a class of lipids that are a major compo- cascade pathway.
nent of all cell membranes. Glycerophospholipids, including Using prostate-specific Pten knockout mice, an immune-
phosphatidylcholine (PC), phosphatidylserine (PS), phos- competent, orthotopic prostate cancer model, and diets with
phatidylethanolamine (PE), and phosphatidylinositol (PI), defined PUFA levels, we found that n-3 fatty acid reduced
contain a diglyceride, a phosphate group, and a simple prostate tumor growth, slowed histopathological progression,
organic molecule, such as choline or serine. and increased survival, whereas n-6 fatty acid had opposite
Dietary PUFA can influence the fatty acid composition effects. Introducing an n-3 desaturase, which converts n-6
of glycerophospholipids in cell membranes. In mammals, the to n-3 fatty acid, into the Pten knockout mice fed an n-6
sn-1 position on the glycerol backbone of glycerophospho- diet reduced tumor growth similarly to mice fed the n-3 diet.
lipids is usually linked to a saturated fatty acid such as stearic Tumors from mice on the n-3 diet had lower proportions of
acid (SA, 18:0), and the sn-2 position is linked to an n-6 PUFA, phosphorylated BAD and higher apoptotic indexes compared
such as AA. Feeding cells in culture, or animals, with n-3 with tumors from mice on the n-6 diet. These data suggest
PUFA can replace n-6 with n-3 fatty acid at the sn-2 position that n-3 PUFA can promote BAD-dependent apoptosis to
of glycerophospholipids; this is considered a diet-induced sn- modulate prostate cancer development [17]. We also found
2 fatty acid moiety change [17, 38, 39]. We have analyzed the that PUFAs modify glycerophospholipid content. DHA can
incorporation efficiency of PUFAs into glycerophospholipids replace the fatty acid at the sn-2 position of the glycerol
in prostate cancer cells. Approximately 25% of input albumin- backbone, thereby changing the species of phospholipid.
conjugated fatty acids were incorporated into cells within 48 DHA also inhibited AKTT308 but not AKTS473 phosphoryla-
hours. The majority of these newly integrated PUFAs were in tion, altered PIP3 and phospho-AKTS473 protein localization,
the form of PC and PE [40]. These data clearly suggest that decreased pPDPK1S241 -AKT and AKT-BAD interaction, and
the fatty acid at the sn-2 position of glycerophospholipids is suppressed prostate tumor growth. Knockdown of BAD
influenced by cellular PUFA uptake. eliminated n-3 PUFA-induced cell death, and reintroduction
n-3 PUFA influences cell membrane conformation and of BAD restored the sensitivity to n-3 fatty acids in vitro.
signaling dynamics. The greater density of n-3 PUFA, com- Knockout of BAD diminished the suppressive effect of n-3
pared to n-6 PUFA, dictates their aggregation near the lipid- PUFA on prostate tumor growth in vivo. These data suggest
water interface. This characteristic can significantly affect that modulation of prostate cancer development by PUFA is
plasma membrane properties, including membrane fluidity, mediated in part through the PI3K/AKT survival pathway
phase behavior, and permeability [41]. These membrane per- [17, 40].
turbations can bring about changes associated with receptor Hu et al. reported that n-3 PUFA-induced apoptosis in
activation, such as diffusional coupling of various peripheral human prostate cancer cells occurs through upregulation
proteins required in G protein-coupled receptor signaling of syndecan-1 (SDC-1) expression followed by concomitant
(GPCR) [42, 43]. Additionally, because of its high level of suppression of PDPK1/AKT/BAD phosphorylation [49]. n-
unsaturation, DHA has very poor affinity for cholesterol, 3 fatty acids may also decrease cell proliferation and induce
which is enriched in lipid rafts of the cell membrane. Lipid apoptotic cell death in human cancer cells by decreasing
rafts are important membrane domains for cell signaling as signal transduction through the AKT/NF𝜅B cell survival
many receptors and proteins are enriched in this domain, pathway and by modulating the PI3K/AKT/p38 MAPK
such as epidermal growth factor receptor (EGFR) [44]. pathway [50, 51].
Hence, incorporation of n-3 PUFA into membrane lipids
can disturb the formation of lipid rafts and suppress raft-
associated cell signal transduction [13, 44]. 4.2. PUFA Mediator. A common fate of unsaturated lipids
The serine/threonine protein kinase AKT (protein kinase released from the membrane is oxidation. AA is the precursor
B) is activated in many solid tumors and hematological of highly bioactive lipid mediators metabolized by a number
malignancies. AKT acts downstream of phosphatidylinositol of enzymes belonging to the cyclooxygenase (COX) and
3-kinase (PI3K) signaling and is a key regulator of multiple lipoxygenase (LOX) families, as well as cytochrome P450.
survival pathways. AKT is known to phosphorylate and COXs have two well-characterized isoforms, COX1 and
inactivate the proapoptotic Bcl-2 family member BAD as one COX2. COX1 is a constitutively expressed gene in most
of its prosurvival tactics [45]. Phosphatidylinositol (PI) is a tissues, whereas COX2 is an immediate-early response gene,
negatively charged constituent of lipid membranes. Specific which is strongly induced in many human malignancies [52].
kinases phosphorylate the hydroxyl groups on positions 3󸀠 , Signal transduction of n-6 PUFA-derived lipids and the effect
4󸀠 , or 5󸀠 of the inositol ring, and PI3K primarily gener- of these lipid mediators on the organism have been well
ates PI-3,4,5-trisphosphate (PIP3 ) from PI-4,5-bisphosphate characterized. For example, AA-derived lipid mediators are
(PI(4,5)P2 ). PIP3 acts as a second messenger to activate pleck- associated with a variety of activities, including inflammation
strin homology (PH) domain-containing proteins, including and cancer. Evidence from human studies also supports the
AKT. Conversely, PIP3 is hydrolyzed to PI(4,5)P2 by PTEN, important role of COXs and LOXs in PUFA metabolism and
4 BioMed Research International

cancer [53–56]. Because of its high expression in inflamma- both Lox12 and Lox15 metabolites are not critical for prostate
tion and cancer, COX2 has been the subject of intense study cancer growth in this model (Chen et al., unpublished).
and proposed as a target for cancer therapy [57–59].
In contrast to n-6 PUFA, the metabolism of n-3 PUFA
is not well understood. Interest in n-3 PUFA-derived lipid 4.3. Fatty Acids Receptors. Lipids are ligands for cell-sur-
mediators began with observations of Greenland Eskimos face G protein-coupled receptors (GPCRs), toll-like recep-
whose diet is rich in marine-derived fish and showed lower tors (TLRs), and peroxisome proliferator-activated receptors
mortality from coronary heart disease and lower preva- (PPARs). G protein-coupled receptors (GPCRs) are impor-
lence of inflammation-related diseases, such as psoriasis, tant signaling molecules for many aspects of cellular function.
inflammatory bowel disease, asthma, rheumatoid arthritis, They are members of a large family that share common
and other autoimmune diseases [60, 61]. Serhan et al. structural motifs, such as seven transmembrane helices and
reported that inflammatory exudates in the murine air pouch the ability to activate heterotrimeric G proteins. Recently,
from mice treated with n-3 PUFA and aspirin contained a several groups reported that unbound free fatty acids can
series of bioactive compounds. Using an unbiased lipidomics activate GPCRs, including GPR40, GPR41, GPR43, GPR84,
approach, they identified and named the EPA-derived E- and GPR120 [3]. Short-chain fatty acids are specific ligands
resolvins (RvE1 and RvE2), DHA-derived D-resolvins (RvD1 for GPR41 and GPR43, medium-chain fatty acids for GPR84,
and RvD2), and (neuro-)protectin (PD1) [62]. and long-chain fatty acids for GPR40 and GPR120 [70–73].
RvE1 and RvE2 are protective in a wide variety of disease Activation of GPR84 receptor by medium-chain fatty acids
models, mainly through their anti-inflammatory activities. triggered the production of the proinflammatory cytokines
RvE1 can resolve inflammation caused by bacterial infection from leukocytes and macrophages. The function of GPR84
of periodontal disease in a rabbit model [63], prevent neo- may be associated with chronic low-grade inflammation-
vascularization after oxygen-induced retinopathy [64], and associated disease [74].
suppress neutrophil infiltration in an acute peritonitis model GPR40 and GPR120 have been reported to be activated
[65]. RvD1, RvD2, and PD1 also have protective activities by long-chain fatty acids such as DHA, EPA, and AA [73, 75].
in a variety of animal models, including models of lung As a G protein-coupled receptor, GPR40 can activate the
injury, insulin resistance, peritonitis, wound healing, and phospholipase C and phosphatidylinositol signaling path-
atherosclerosis [66]. RvD1 was shown to reduce leukocyte ways [76]. Although GPR40 is preferentially expressed in
infiltration in murine inflammatory exudates [67] and RvD2 pancreatic 𝛽-cells and is known to mediate insulin secretion
was shown to reverse inflammatory pain in mice [68]. PD1 [77], several groups showed that it is expressed in the
has been reported to regulate amyloid beta secretion and brain where it mediates the antinociceptive activity of DHA
thereby improve neuronal survival in a mouse model of [78, 79]. Recently, Oh and others reported that GPR120
Alzheimer’s disease [69]. Although ample data indicate that functions as an n-3 PUFA receptor in vitro and in vivo [3]
these n-3 PUFA-derived mediators can resolve inflammation, and suggested that diminished activation of GPR120 can be
little is known about their role in inflammation-related an important contributor to obesity, insulin resistance, and
cancers, such as prostate and colon cancers. tissue inflammation [80, 81]. GPR120 is highly expressed
Due to the existence of multiple oxygenases, the role in adipose tissue, proinflammatory bone marrow-derived
of each enzyme in the development of prostate cancer has CD11C+ macrophages (BMDCs), mature adipocytes, and
not been studied systematically in a single system or animal monocytic RAW 264.7 macrophage cells. DHA strongly
model. Furthermore, studies performed in animals rarely inhibited LPS-induced phosphorylation of JNK and IKK𝛽,
take diet into account. To systematically assess the interaction I𝜅B degradation, cytokine secretion- and inflammatory gene
between oxygenases and dietary PUFA in a single animal expression level in GPR120-positive cells. These effects of
model of prostate cancer, we knocked out Cox1, Cox2, DHA were completely prevented by GPR120 knockdown,
Lox5, Lox12, or Lox15 in prostate-specific Pten null mice. demonstrating that these anti-inflammatory effects were
specifically exerted through GPR120. An n-3 PUFA diet
Our preliminary results indicate that tumor growth was
containing 27% fish oil led to improved insulin sensitivity
significantly increased in Cox1−/− Pten null mice on n-3 diet with increased glucose infusion rates, enhanced muscle
compared to Cox1-wild-type Pten null littermates. This result insulin sensitivity, and greater hepatic insulin sensitivity.
suggests that Cox1 is required for the protective effects of The n-3 PUFA diet had no effect in the GPR120 knockout
n-3 PUFA. Interestingly, tumor growth was decreased in n-6 (KO) mice. On chow diets, the GPR120 KO mice showed
PUFA fed Cox1−/− Pten-null mice compared to n-6 fed Cox1- moderate insulin resistance with no changes in food intake or
wildtype Pten-null mice, suggesting that n-6 metabolites of body weight. On high-fat diet (HFD), the GPR120 KO mice
Cox1 promote tumor growth. Loss of Cox2 reduced prostate gained more weight than wild-type controls [3]. In humans,
tumor growth on both n-3 and n-6 diets, suggesting that GPR120 expression in adipose tissue is significantly higher in
the suppressive effect of n-3 PUFA is independent of Cox2 obese individuals than in lean controls [80]. Ichimura and
metabolism. Loss of Lox5 reduced prostate tumor growth colleagues compared sequences of GPR120 exons in obese
on n-6 diet but had no effect on n-3 diet; loss of Lox12 or populations and discovered a deleterious nonsynonymous
Lox15 did not affect prostate tumor growth on either diet. mutation (R270H) [80]. Their population study showed that
These results suggest that the promotion of prostate tumor the GPR120R270H variant correlated with obesity. Further
growth by n-6 diet is dependent on Lox-5 metabolism and investigation in vitro demonstrated that the GPR120R270H
BioMed Research International 5

variant was unable to respond to long-chain fatty acid reduction in tumor growth as a result of dietary n-3 PUFA
stimulation. This inactive mutant of GPR120 may contribute is accompanied by an increase in the expression of secreted
to its significant association with obesity [80]. SDC-1 [49, 101].
Toll-like receptors (TLRs) are transmembrane glycopro- Several other receptors have been suggested as targets for
tein receptors that are important regulators of the innate n-3 PUFA action. Turk et al. reported that DHA can induce
immune system. TLRs are considered a link between innate the alteration in both the lateral and subcellular localization
(nonspecific) and adaptive (specific) immunity and con- of EGFR and suppress EGFR signaling, which suggests
tribute to the immune system’s capacity to efficiently combat implications for the molecular basis of cancer prevention by
pathogens [82]. TLR expression is increased in tumors, DHA [44, 102]. It is also reported that DHA can increase
including breast, colorectal, melanoma, lung, prostate, pan- CD95 (Fas ligand death receptor) cell surface expression
creatic, and liver cancer [83]. Activation of TLRs triggers and may mediate CD95-induced apoptosis [103]. For more
a signaling cascade producing inflammatory cytokines that information about PUFA receptor interaction, please refer to
recruit components of the adaptive immune system to kill the a review by Lee et al. [104].
pathogen [84]. Among the family of TLRs, TLR4 and TLR9
have been reported to be associated with prostate cancer
[85–87]. Panigrahy et al. reported that saturated fatty acids 4.4. Other Mechanisms. Nuclear factor erythroid-2-related
activated, and DHA inhibited, TLR2- and TLR4-mediated factor 2 (Nrf2) is a basic leucine zipper transcription factor.
proinflammatory activity in a cell culture system [88]. Satu- Nrf2 is sequestered in the cytoplasm by Kelch-like ECH-
rated fatty acids may stimulate the TLR4 signaling pathway to associated protein 1 (Keap1) under basal conditions. When
trigger the production of proinflammatory mediators, which the cell is challenged by oxidative stress, Nrf2 is released from
may contribute to neuronal death [89]. Keap1 inhibition, translocates to the nucleus, forms a com-
PPARs (PPAR𝛼, PPAR𝛽/𝛿, and PPAR𝛾) are a super- plex with other factors, and activates transcription of genes
family of ligand-activated transcription factors and nuclear containing an antioxidant response element (ARE) in their
hormone receptors. The syndecan family of cell surface promoter region. Nrf2 has been reported to play an important
proteoglycans share a structure of small, conserved cyto- role in lung injury reversal, human endothelial cell survival,
plasmic and transmembrane domains and larger, distinct neuroinflammation, hyperoxia, lung damage from cigarette
ectodomain. They are implicated in a variety of physiologic smoking, and impaired function of macrophages [105]. Other
and pathologic processes such as nutrient metabolism, energy studies suggest that Nrf2 suppresses inflammation by inhibit-
homeostasis, inflammation, and cancer. Growing evidence ing NF𝜅B activation through regulation of redox balance,
has demonstrated that PPAR𝛾 serves as a tumor suppressor calcium signaling, and PPARs [106]. Various human cancers,
in cancer (see review of Robbins and Nie, 2012) [90]. n-3 such as lung cancer, frequently exhibit increased levels of
PUFA can induce apoptosis in human prostate cancer cells Nrf2 [107]. Downregulation of nuclear Nrf2 gene expression
by activating the nuclear receptor PPAR𝛾 and upregulating by RNAi-mediated silencing in nonsmall cell lung cancer
the PPAR𝛾 target gene, syndecan-1 (SDC-1) [49, 91]. It has inhibits tumor growth and increases efficacy of chemotherapy
been suggested that n-3 PUFA induces cell apoptosis in [108]. Cancer cells are suspected to hijack the Keap1-Nrf2
prostate cancer via a mechanism of LOX15-mediated SDC- system as a means to acquire malignant properties. Indeed,
1-dependent suppression of PDPK1/AKT/BAD phosphory- the prognosis of patients carrying Nrf2-positve cancers is
lation [49, 91]. SDC-1 upregulation by DHA has also been poor [105]. Oxidized n-3 fatty acids can react directly with the
demonstrated in human breast cancer cells [19, 92–94] and in negative regulator of Nrf2, Keap1, by dissociating them and
n-3 PUFA-enriched mammary glands and liver of fat-1 mice inducing Nrf2-directed gene expression [109]. For example,
[95]. n-3 PUFA mediators can activate Nrf2 in vascular endothelial
SDC-1 plays several important cellular functions. It reg- cells to prevent oxidative stress-induced cytotoxicity [110].
ulates many steps of leukocyte recruitment in noninfectious DHA and EPA can induce Nrf2 expression and suppress
inflammatory diseases, attenuates inflammation by modu- lipopolysaccharide-(LPS-) induced inflammation [111]. Evi-
lating heparin sulfate-binding proinflammatory factors, and dence of Nrf2-mediated response modulated by n-3 PUFA
plays a key role in the normal remodeling of injured cardiac in prostate cancer came from a randomized clinical trial.
tissues [96]. Loss of cell surface SDC-1, seen in many Eighty-four men with low-risk prostate cancer were stratified
carcinomas such as skin cancer and colorectal adenocar- based on self-reported dietary consumption of fish oil.
cinomas, favors acquisition of the metastatic phenotype in Exploratory pathway analyses of rank-ordered genes revealed
cancer cells [97]. There is very limited information about the modulation of Nrf2 or Nrf2-mediated oxidative response
SDC-1 expression in prostate cancer. Some studies reported after 3 months of fish oil supplementation (𝑃 = 0.01) [112].
an inverse relationship between SDC-1 and Gleason score Calcium (Ca2+ ) signaling is a ubiquitous mechanism in
[98, 99], but a tissue microarray analysis in another study the control of cell function. The transient receptor potential
showed an increase in SDC-1 with tumor progression [100]. channels (TRP channels) are 6 transmembrane-spanning
Our own studies have shown reduced expression of SDC- proteins with both amino and carboxyl tails located on
1 in prostate cancer cell lines compared to normal prostate the intracellular side of the membrane. Ca2+ flux through
epithelial cells and lower expression in androgen-dependent TRP channels located in the plasma membrane and in the
LNCaP cells compared to androgen-independent PC3 and membranes of excitable intracellular organelles can promote
DU145 cells [49]. In a mouse prostate cancer model, the changes in intracellular free Ca2+ concentrations and the
6 BioMed Research International

Table 1: Mechanism of n-3 PUFA action.


Target Mechanism of action References
Membrane PIPs Suppress PI3K/AKT/BAD survival pathway [17, 40]
SDC-1 Suppress PI3K/AKT/BAD and AKT/NF𝜅B survival pathway [49]
Unknown Suppress AKT/NF𝜅B or PI3K/AKT/p38 MAPK pathway [50, 51]
GPR40 Mediate insulin secretion, cell proliferation [75–77]
GPR120 Anti-inflammation [3, 80]
TLRs Suppress the production of proinflammatory mediators [88, 89]
PPARg Induce cell apoptosis [90]
Nrf2 Suppress inflammation, reduce oxidative stress [106, 109, 110]
Calcium signaling Inhibit cell proliferation [114]
COXs Promote inflammation [52]
LOXs Promote inflammation [53]
E-resolvins Resolve inflammation [62]
D-resolvins Resolve inflammation [62]

membrane potential, which can modulate the driving force metabolic processes, including 𝛽-oxidation, lipid release
for other ions and Ca2+ itself [113]. Evidence suggests that from glycerophospholipids, cellular signaling of membrane
TRP channel function can be modulated both directly and bound proteins, eicosanoid synthesis, and direct activation
indirectly by n-3 fatty acids [114]. It was demonstrated that of nuclear receptors and gene transcription, all of which
DHA and EPA at physiological concentrations have the ability may influence the development and progression of prostate
to evoke small currents, which seems to be dependent on cancer. Overall, there seems to be an exceptionally broad
the previous sensitization of the channel by protein kinase potential for the mechanisms mediating cancer prevention
C (PKC). Whether these effects are due to the action of by n-3 PUFA (summarized in Table 1). We expect that new
these PUFAs on the agonist binding site or are due to research in lipidomics and metabolomics will provide new
conformational changes caused by TRP protein interactions techniques and approaches to answering the many questions
with the lipid bilayer requires further investigation [115]. that remain regarding the mechanisms underlying the health
Recent findings also indicate that TRP channel function benefits of n-3 PUFA.
can be modulated by D and E resolvins. Resolvin binding
on GPCRs seems to be part of the mechanism underlying
the resolvin-mediated regulation of TRP channel function
[116]. DHA significantly reduces oxidative stress-induced Abbreviations
endothelial cell Ca2+ influx. This effect might be associated, PUFA: Polyunsaturated fatty acid
at least in part, with altered lipid composition in membrane EPA: Eicosapentaenoic acid (20:5, n-3)
caveolar rafts [117]. Ca2+ has been shown to be essential DHA: Docosahexaenoic acid (22:6, n-3)
for increased cell proliferation in prostate cells [118]. Sun LA: Linoleic acid (18:2, n-6)
et al. observed significantly higher Ca2+ influx in prostate AA: Arachidonic acid (20:4, n-6)
cancer cells. They reported that high ratio of Ca2+ /Mg2+ ALA: Alpha linolenic acid (18:3, n-3)
facilitated Ca2+ influx and led to a significant increase in FADS: Fatty acid desaturase
cell proliferation of prostate cancer [119]. Thus, one could GLA: Gamma-linolenic acid (18:3, n-6)
speculate that n-3 PUFA might indirectly modulate prostate DGLA: Dihomo-gamma-linolenic acid (20:3, n-6)
cancer growth by directly modulating Ca2+ influx. KAR: 3-Ketoacyl-CoA reductase
HACD: 3-Hydroxyacyl-CoA dehydratase
TECR: Trans-2,3-enoyl-CoA reductase
5. Conclusions DPA: Docosapentaenoic acid (22:5, n-3)
PC: Phosphatidylcholine
Cancer incidence and mortality are high in the Western world PS: Phosphatidylserine
and a high n-6 to n-3 PUFA ratio in the Western diet may PE: Phosphatidylethanolamine
be a contributing factor. There is much evidence to suggest PI: Phosphatidylinositol
that n-3 PUFA has antiproliferative effects in cancer cell SA: Stearic acid (18:0)
lines, animal models, and humans. Direct effects on cancer GPCR: G Protein-coupled receptor
cells and indirect effects on the host immune system (anti- EGFR: Epidermal growth factor receptor
inflammation) likely contribute to the inhibitory effect of n-3 AKT: Serine/threonine protein kinase (protein kinase B)
fatty acids on tumor growth; however, further investigation PI3K: Phosphatidylinositol-3-kinase
is warranted. n-3 PUFA may also regulate other complex PIP3 : PI-3,4,5-trisphosphate
BioMed Research International 7

PH: Pleckstrin homology Reproduction Nutrition Development, vol. 45, no. 5, pp. 581–597,
PDPK1: Phosphoinositide-dependent kinase-1 2005.
SDC-1: Syndecan-1 [13] C. M. Williams and G. Burdge, “Long-chain n-3 PUFA: plant v.
COX: Cyclooxygenase marine sources,” Proceedings of the Nutrition Society, vol. 65, no.
LOX: Lipoxygenase 1, pp. 42–50, 2006.
RvE1: E-Resolvin1 [14] R. J. Pawlosky, J. R. Hibbeln, J. A. Novotny, and N. Salem, “Physi-
RvD1: D-Resolvin1 ological compartmental analysis of 𝛼-linolenic acid metabolism
TLR: Toll-like receptor in adult humans,” Journal of Lipid Research, vol. 42, no. 8, pp.
PPAR: Peroxisome proliferator-activated receptor 1257–1265, 2001.
MCFA: Medium-chain fatty acid [15] A. P. Simopoulos, “Evolutionary aspects of omega-3 fatty acids
BMDC: Bone marrow-derived CD11C+ macrophage in the food supply,” Prostaglandins Leukotrienes and Essential
KO: Knockout Fatty Acids, vol. 60, no. 5-6, pp. 421–429, 1999.
HFD: High-fat diet [16] A. P. Simopoulos, “The importance of the ratio of omega-
Nrf2: Nuclear factor erythroid-2-related factor 2 6/omega-3 essential fatty acids,” Biomedicine and Pharma-
cotherapy, vol. 56, no. 8, pp. 365–379, 2002.
Keap1: Kelch-like ECH-associated protein 1
ARE: Antioxidant response element [17] I. M. Berquin, Y. Min, R. Wu et al., “Modulation of prostate
cancer genetic risk by omega-3 and omega-6 fatty acids,” Journal
LPS: Lipopolysaccharide
of Clinical Investigation, vol. 117, no. 7, pp. 1866–1875, 2007.
PKC: Protein kinase C.
[18] G. Calviello, F. Resci, S. Serini et al., “Docosahexaenoic
acid induces proteasome-dependent degradation of 𝛽-catenin,
References down-regulation of survivin and apoptosis in human colorectal
cancer cells not expressing COX-2,” Carcinogenesis, vol. 28, no.
[1] H. O. Bang, J. Dyerberg, and A. B. Nielsen, “Plasma lipid and 6, pp. 1202–1209, 2007.
lipoprotein pattern in Greenlandic West-coast Eskimos,” The [19] H. Sun, I. M. Berquin, R. T. Owens, J. T. O’Flaherty, and
Lancet, vol. 1, no. 7710, pp. 1143–1145, 1971. I. J. Edwards, “Peroxisome proliferator-activated receptor 𝛾-
[2] J. Dyerberg, H. O. Bang, and N. Hjorne, “Fatty acid composition mediated up-regulation of syndecan-1 by n-3 fatty acids pro-
of the plasma lipids in Greenland Eskimos,” The American motes apoptosis of human breast cancer cells,” Cancer Research,
Journal of Clinical Nutrition, vol. 28, no. 9, pp. 958–966, 1975. vol. 68, no. 8, pp. 2912–2919, 2008.
[3] D. Y. Oh, S. Talukdar, E. J. Bae et al., “GPR120 is an Omega- [20] I. M. Berquin, I. J. Edwards, and Y. Q. Chen, “Multi-targeted
3 fatty acid receptor mediating potent anti-inflammatory and therapy of cancer by omega-3 fatty acids,” Cancer Letters, vol.
insulin-sensitizing effects,” Cell, vol. 142, no. 5, pp. 687–698, 269, no. 2, pp. 363–377, 2008.
2010. [21] Y. Q. Chen, I. J. Edwards, S. J. Kridel, T. Thornburg, and I.
[4] A. Ariel and C. N. Serhan, “Resolvins and protectins in M. Berquin, “Dietary fat-gene interactions in cancer,” Cancer
the termination program of acute inflammation,” Trends in Metastasis Reviews, vol. 26, pp. 535–551, 2007.
Immunology, vol. 28, no. 4, pp. 176–183, 2007. [22] S. Wang, J. Wu, J. Suburu, Z. Gu, J. Cai et al., “Effect of dietary
[5] C. A. Daley, A. Abbott, P. S. Doyle, G. A. Nader, and S. Larson, polyunsaturated fatty acids on castration-resistant Pten-null
“A review of fatty acid profiles and antioxidant content in grass- prostate cancer,” Carcinogenesis, vol. 33, pp. 404–412, 2012.
fed and grain-fed beef,” Nutrition Journal, vol. 9, no. 1, article 10, [23] K. K. Carroll, “Dietary fat and cancer: specific action or caloric
2010. effect?” Journal of Nutrition, vol. 116, no. 6, pp. 1130–1132, 1986.
[6] J. T. Bernert Jr. and H. Sprecher, “Studies to determine the role [24] M. Gerber, “Omega-3 fatty acids and cancers: a systematic
rates of chain elongation and desaturation play in regulating update review of epidemiological studies,” British Journal of
the unsaturated fatty acid composition of rat liver lipids,” Nutrition, vol. 107, Supplement 2, pp. 228–239, 2012.
Biochimica et Biophysica Acta, vol. 398, no. 3, pp. 354–363, 1975.
[25] P. D. Terry, T. E. Rohan, and A. Wolk, “Intakes of fish and marine
[7] I. M. Berquin, I. J. Edwards, S. J. Kridel, and Y. Q. Chen, fatty acids and the risks of cancers of the breast and prostate and
“Polyunsaturated fatty acid metabolism in prostate cancer,” of other hormone-related cancers: a review of the epidemiologic
Cancer and Metastasis Reviews, vol. 30, no. 3-4, pp. 295–309, evidence,” The American Journal of Clinical Nutrition, vol. 77, no.
2011. 3, pp. 532–543, 2003.
[8] J. V. Swinnen, K. Brusselmans, and G. Verhoeven, “Increased [26] C. H. MacLean, S. J. Newberry, W. A. Mojica et al., “Effects of
lipogenesis in cancer cells: new players, novel targets,” Current omega-3 fatty acids on cancer risk: a systematic review,” Journal
Opinion in Clinical Nutrition and Metabolic Care, vol. 9, no. 4, of the American Medical Association, vol. 295, no. 4, pp. 403–415,
pp. 358–365, 2006. 2006.
[9] F. P. Kuhajda, “Fatty acid synthase and cancer: new application [27] Y. Q. Chen, I. M. Berquin, L. W. Daniel et al., “Omega-3
of an old pathway,” Cancer Research, vol. 66, no. 12, pp. 5977– fatty acids and cancer risk,” Journal of the American Medical
5980, 2006. Association, vol. 296, no. 3, pp. 278–282, 2006.
[10] J. A. Menendez and R. Lupu, “Fatty acid synthase and the [28] M. C. Schumacher, B. Laven, F. Petersson et al., “A comparative
lipogenic phenotype in cancer pathogenesis,” Nature Reviews study of tissue 𝜔-6 and 𝜔-3 polyunsaturated fatty acids (PUFA)
Cancer, vol. 7, no. 10, pp. 763–777, 2007. in benign and malignant pathologic stage pT2a radical prostate-
[11] J. Suburu and Y. Q. Chen, “Lipids and prostate cancer,” ctomy specimens,” Urologic Oncology, vol. 31, no. 3, pp. 318–324,
Prostaglandins and Other Lipid Mediators, vol. 98, pp. 1–10, 2012. 2013.
[12] G. C. Burdge and P. C. Calder, “Conversion of 𝛼-linolenic acid [29] C. D. Williams, B. M. Whitley, C. Hoyo, D. J. Grant, J. D. Iraggi et
to longer-chain polyunsaturated fatty acids in human adults,” al., “A high ratio of dietary n-6/n-3 polyunsaturated fatty acids
8 BioMed Research International

is associated with increased risk of prostate cancer,” Nutrition [45] S. R. Datta, H. Dudek, T. Xu et al., “Akt phosphorylation of BAD
Research, vol. 31, no. 1, pp. 1–8, 2011. couples survival signals to the cell- intrinsic death machinery,”
[30] M. D. Brown, C. Hart, E. Gazi, P. Gardner, N. Lockyer, and N. Cell, vol. 91, no. 2, pp. 231–241, 1997.
Clarke, “Influence of omega-6 PUFA arachidonic acid and bone [46] T. Maehama and J. E. Dixon, “The tumor suppressor,
marrow adipocytes on metastatic spread from prostate cancer,” PTEN/MMAC1, dephosphorylates the lipid second messenger,
British Journal of Cancer, vol. 102, no. 2, pp. 403–413, 2010. phosphatidylinositol 3,4,5-trisphosphate,” Journal of Biological
[31] C. R. Ritch, R. L. Wan, L. B. Stephens et al., “Dietary fatty Chemistry, vol. 273, no. 22, pp. 13375–13378, 1998.
acids correlate with prostate cancer biopsy grade and volume [47] J. A. Engelman, J. Luo, and L. C. Cantley, “The evolution
in Jamaican men,” Journal of Urology, vol. 177, no. 1, pp. 97–101, of phosphatidylinositol 3-kinases as regulators of growth and
2007. metabolism,” Nature Reviews Genetics, vol. 7, no. 8, pp. 606–619,
[32] N. Sawada, M. Inoue, M. Iwasaki, S. Sasazuki, T. Shimazu et 2006.
al., “Consumption of n-3 fatty acids and fish reduces risk of [48] T. F. Franke, “PI3K/Akt: getting it right matters,” Oncogene, vol.
hepatocellular carcinoma,” Gastroenterology, vol. 142, pp. 1468– 27, no. 50, pp. 6473–6488, 2008.
1475, 2012. [49] Y. Hu, H. Sun, R. T. Owens et al., “Syndecan-1-dependent
[33] S. Sasazuki, M. Inoue, M. Iwasaki et al., “Intake of n-3 and suppression of PDK1/Akt/Bad signaling by docosahexaenoic
n-6 polyunsaturated fatty acids and development of colorectal acid induces apoptosis in prostate cancer,” Neoplasia, vol. 12, no.
cancer by subsite: Japan Public Health Center-based prospective 10, pp. 826–836, 2010.
study,” International Journal of Cancer, vol. 129, no. 7, pp. 1718– [50] P. D. Schley, H. B. Jijon, L. E. Robinson, and C. J. Field,
1729, 2011. “Mechanisms of omega-3 fatty acid-induced growth inhibition
[34] J. E. Chavarro, M. J. Stampfer, H. Li, H. Campos, T. Kurth, in MDA-MB-231 human breast cancer cells,” Breast Cancer
and J. Ma, “A prospective study of polyunsaturated fatty acid Research and Treatment, vol. 92, no. 2, pp. 187–195, 2005.
levels in blood and prostate cancer risk,” Cancer Epidemiology [51] J. L. D. Toit-Kohn, L. Louw, and A. M. Engelbrecht, “Docosa-
Biomarkers and Prevention, vol. 16, no. 7, pp. 1364–1370, 2007. hexaenoic acid induces apoptosis in colorectal carcinoma cells
[35] K. M. Szymanski, D. C. Wheeler, and L. A. Mucci, “Fish by modulating the PI3 kinase and p38 MAPK pathways,” Journal
consumption and prostate cancer risk: a review and meta- of Nutritional Biochemistry, vol. 20, no. 2, pp. 106–114, 2009.
analysis,” The American Journal of Clinical Nutrition, vol. 92, no. [52] R. N. DuBois, S. B. Abramson, L. Crofford et al., “Cyclooxyge-
5, pp. 1223–1233, 2010. nase in biology and disease,” FASEB Journal, vol. 12, no. 12, pp.
[36] N. Kobayashi, R. J. Barnard, S. M. Henning et al., “Effect of 1063–1073, 1998.
altering dietary 𝜔-6/𝜔-3 fatty acid ratios on prostate cancer [53] G. P. Pidgeon, J. Lysaght, S. Krishnamoorthy et al., “Lipoxy-
membrane composition, cyclooxygenase-2, and prostaglandin genase metabolism: roles in tumor progression and survival,”
E 2,” Clinical Cancer Research, vol. 12, no. 15, pp. 4662–4670, Cancer and Metastasis Reviews, vol. 26, no. 3-4, pp. 503–524,
2006. 2007.
[37] M. D. Brown, C. A. Hart, E. Gazi, S. Bagley, and N. W. Clarke, [54] Y. N. Ye, W. K. K. Wu, V. Y. Shin, I. C. Bruce, B. C. Y.
“Promotion of prostatic metastatic migration towards human Wong, and C. H. Cho, “Dual inhibition of 5-LOX and COX-
bone marrow stoma by Omega 6 and its inhibition by Omega 2 suppresses colon cancer formation promoted by cigarette
3 PUFAs,” British Journal of Cancer, vol. 94, no. 6, pp. 842–853, smoke,” Carcinogenesis, vol. 26, no. 4, pp. 827–834, 2005.
2006. [55] B. K. Sharma, P. Pilania, and P. Singh, “Modeling of
[38] M. C. Chabot, J. D. Schmitt, B. C. Bullock, and R. L. Wykle, cyclooxygenase-2 and 5-lipooxygenase inhibitory activity
“Reacylation of platelet activating factor with eicosapentaenoic of apoptosis-inducing agents potentially useful in prostate
acid in fish-oil-enriched monkey neutrophils,” Biochimica et cancer chemotherapy: derivatives of diarylpyrazole,” Journal of
Biophysica Acta, vol. 922, no. 2, pp. 214–220, 1987. Enzyme Inhibition and Medicinal Chemistry, vol. 24, no. 2, pp.
[39] M. Picq, P. Chen, M. Perez, M. Michaud, E. Vericel et al., “DHA 607–615, 2009.
metabolism: targeting the brain and lipoxygenation,” Molecular [56] M. Diederich, C. Sobolewski, C. Cerella, M. Dicato, and L.
Neurobiology, vol. 42, pp. 48–51, 2010. Ghibelli, “The role of cyclooxygenase-2 in cell proliferation and
[40] Z. Gu, J. Wu, S. Wang et al., “Polyunsaturated fatty acids affect cell death in human malignancies,” International Journal of Cell
the localization and signaling of PIP3/AKT in prostate cancer Biology, Article ID 215158, 2010.
cells,” Carcinogenesis, 2013. [57] M. T. Wang, K. V. Honn, and D. Nie, “Cyclooxygenases,
[41] S. R. Wassall and W. Stillwell, “Docosahexaenoic acid domains: prostanoids, and tumor progression,” Cancer and Metastasis
the ultimate non-raft membrane domain,” Chemistry and Reviews, vol. 26, no. 3-4, pp. 525–534, 2007.
Physics of Lipids, vol. 153, no. 1, pp. 57–63, 2008. [58] D. G. Menter, R. L. Schilsky, and R. N. DuBois,
[42] N. V. Eldho, S. E. Feller, S. Tristram-Nagle, I. V. Polozov, and “Cyclooxygenase-2 and cancer treatment: understanding
K. Gawrisch, “Polyunsaturated docosahexaenoic vs docosapen- the risk should be worth the reward,” Clinical Cancer Research,
taenoic acid - Differences in lipid matrix properties from the vol. 16, no. 5, pp. 1384–1390, 2010.
loss of one double bond,” Journal of the American Chemical [59] A. C. Reese, V. Fradet, and J. S. Witte, “𝜔-3 Fatty acids, genetic
Society, vol. 125, no. 21, pp. 6409–6421, 2003. variants in COX-2 and prostate cancer,” Journal of Nutrigenetics
[43] D. C. Mitchell, S. L. Niu, and B. J. Litman, “Quantifying the and Nutrigenomics, vol. 2, no. 3, pp. 149–158, 2009.
differential effects of DHA and DPA on the early events in visual [60] H. O. Bang, J. Dyerberg, and N. Hjorne, “The composition of
signal transduction,” Chemistry and Physics of Lipids, vol. 165, food consumed by Greenland Eskimos,” Acta Medica Scandi-
pp. 393–400, 2012. navica, vol. 200, no. 1-2, pp. 69–73, 1976.
[44] H. F. Turk, R. Barhoumi, and R. S. Chapkin, “Alteration of EGFR [61] J. Dyerberg, H. O. Bang, E. Stoffersen, S. Moncada, and J. R.
spatiotemporal dynamics suppresses signal transduction,” PLoS Vane, “Eicosapentaenoic acid and prevention of thrombosis and
One, vol. 7, no. 6, Article ID e39682, 2012. atherosclerosis?” The Lancet, vol. 2, no. 8081, pp. 117–119, 1978.
BioMed Research International 9

[62] C. N. Serhan, C. B. Clish, J. Brannon, S. P. Colgan, N. Chiang, [77] T. Alquier, M. L. Peyot, M. G. Latour et al., “Deletion of
and K. Gronert, “Novel functional sets of lipid-derived medi- GPR40 impairs glucose-induced insulin secretion in vivo in
ators with antiinflammatory actions generated from omega-3 mice without affecting intracellular fuel metabolism in islets,”
fatty acids via cyclooxygenase 2-nonsteroidal antiinflammatory Diabetes, vol. 58, no. 11, pp. 2607–2615, 2009.
drugs and transcellular processing,” Journal of Experimental [78] K. Nakamoto, T. Nishinaka, K. Matsumoto, F. Kasuya, M.
Medicine, vol. 192, no. 8, pp. 1197–1204, 2000. Mankura et al., “Involvement of the long-chain fatty acid
[63] H. Hasturk, A. Kantarci, T. Ohira et al., “RvE1 protects from receptor GPR40 as a novel pain regulatory system,” Brain
local inflammation and osteoclast-mediated bone destruction Research, vol. 1432, pp. 74–83, 2012.
in periodontitis,” FASEB Journal, vol. 20, no. 2, pp. 401–403, [79] D. Ma, M. Zhang, C. P. Larsen et al., “DHA promotes the
2006. neuronal differentiation of rat neural stem cells transfected with
[64] K. M. Connor, J. P. Sangiovanni, C. Lofqvist et al., “Increased GPR40 gene,” Brain Research, vol. 1330, pp. 1–8, 2010.
dietary intake of 𝜔-3-polyunsaturated fatty acids reduces patho- [80] A. Ichimura, A. Hirasawa, O. Poulain-Godefroy, A. Bonnefond,
logical retinal angiogenesis,” Nature Medicine, vol. 13, no. 7, pp. T. Hara et al., “Dysfunction of lipid sensor GPR120 leads to
868–873, 2007. obesity in both mouse and human,” Nature, vol. 483, pp. 350–
[65] S. Ogawa, D. Urabe, Y. Yokokura, H. Arai, M. Arita, and M. 354, 2012.
Inoue, “Total synthesis and bioactivity of resolvin E2,” Organic [81] D. Y. Oh and J. M. Olefsky, “Omega 3 fatty acids and GPR120,”
Letters, vol. 11, no. 16, pp. 3602–3605, 2009. Cell Metabolism, vol. 15, pp. 564–565, 2012.
[66] K. H. Weylandt, C. Y. Chiu, B. Gomolka, S. F. Waechter, and B. [82] T. Kawai and S. Akira, “TLR signaling,” Cell Death and Differ-
Wiedenmann, “Omega-3 fatty acids and their lipid mediators: entiation, vol. 13, no. 5, pp. 816–825, 2006.
towards an understanding of resolvin and protectin formation,” [83] L. A. Ridnour, R. Y. Cheng, C. H. Switzer, J. L. Heinecke, S.
Prostaglandins and Other Lipid Mediators, vol. 97, pp. 73–82, Ambs et al., “Molecular Pathways: toll-like receptors in the
2012. tumor microenvironment: poor prognosis or new therapeutic
[67] Y. P. Sun, S. F. Oh, J. Uddin et al., “Resolvin D1 and its opportunity,” Clinical Cancer Research, vol. 19, no. 6, pp. 1340–
aspirin-triggered 17R epimer: stereochemical assignments, anti- 1346, 2012.
inflammatory properties, and enzymatic inactivation,” Journal [84] J. Krishnan, G. Lee, and S. Choi, “Drugs targeting toll-like
of Biological Chemistry, vol. 282, no. 13, pp. 9323–9334, 2007. receptors,” Archives of Pharmacal Research, vol. 32, no. 11, pp.
[68] C. K. Park, Z. Z. Xu, T. Liu, N. Lu, C. N. Serhan et al., “Resolvin 1485–1502, 2009.
D2 is a potent endogenous inhibitor for transient receptor [85] A. Paone, D. Starace, R. Galli et al., “Toll-like receptor 3 triggers
potential subtype V1/A1, inflammatory pain, and spinal cord apoptosis of human prostate cancer cells through a PKC-𝛼-
synaptic plasticity in mice: distinct roles of resolvin D1, D2, and dependent mechanism,” Carcinogenesis, vol. 29, no. 7, pp. 1334–
E1,” Journal of Neuroscience, vol. 31, pp. 18433–18438, 2011. 1342, 2008.
[69] Y. Zhao, F. Calon, C. Julien et al., “Docosahexaenoic acid- [86] S. L. Zheng, K. Augustsson-Bälter, B. Chang et al., “Sequence
derived neuroprotectin D1 induces neuronal survival via variants of toll-like receptor 4 are associated with prostate
secretase- and PPAR𝛾-mediated mechanisms in Alzheimer’s cancer risk: results from the cancer prostate in Sweden study,”
disease models,” PLoS ONE, vol. 6, no. 1, Article ID e15816, 2011. Cancer Research, vol. 64, no. 8, pp. 2918–2922, 2004.
[87] M. R. Väisänen, T. Väisänen, A. Jukkola-Vuorinen et al.,
[70] H. Tazoe, Y. Otomo, I. Kaji, R. Tanaka, S. I. Karaki, and A.
“Expression of toll-like receptor-9 is increased in poorly differ-
Kuwahara, “Roles of short-chain fatty acids receptors, GPR41
entiated prostate tumors,” Prostate, vol. 70, no. 8, pp. 817–824,
and GPR43 on colonic functions,” Journal of Physiology and
2010.
Pharmacology, vol. 59, supplement 2, pp. 251–262, 2008.
[88] D. Panigrahy, A. Kaipainen, E. R. Greene, and S. Huang,
[71] J. Wang, X. Wu, N. Simonavicius, H. Tian, and L. Ling,
“Cytochrome P450-derived eicosanoids: the neglected pathway
“Medium-chain fatty acids as ligands for orphan G protein-
in cancer,” Cancer Metastasis Reviews, vol. 29, no. 4, pp. 723–735,
coupled receptor GPR84,” Journal of Biological Chemistry, vol.
2010.
281, no. 45, pp. 34457–34464, 2006.
[89] Z. Wang, D. Liu, F. Wang, S. Liu, S. Zhao et al., “Saturated
[72] Y. Itoh, Y. Kawamata, M. Harada et al., “Free fatty acids regulate fatty acids activate microglia via Toll-like receptor 4/NF-kappaB
insulin secretion from pancreatic 𝛽 cells through GPR40,” signalling,” British Journal of Nutrition, vol. 107, pp. 229–241,
Nature, vol. 422, no. 6928, pp. 173–176, 2003. 2012.
[73] A. Hirasawa, K. Tsumaya, T. Awaji et al., “Free fatty acids [90] G. T. Robbins and D. Nie, “PPAR gamma, bioactive lipids, and
regulate gut incretin glucagon-like peptide-1 secretion through cancer progression,” Frontiers in Bioscience, vol. 17, pp. 1816–
GPR120,” Nature Medicine, vol. 11, no. 1, pp. 90–94, 2005. 1834, 2012.
[74] M. Suzuki, S. Takaishi, M. Nagasaki, Y. Onozawa, I. Iino et [91] H. J. Murff, X. O. Shu, H. Li et al., “Dietary polyunsaturated fatty
al., “Medium-chain fatty acid-sensing receptor, GPR84, is a acids and breast cancer risk in Chinese women: a prospective
proinflammatory receptor,” Journal of Biological Chemistry, vol. cohort study,” International Journal of Cancer, vol. 128, no. 6, pp.
288, pp. 10684–10691, 2013. 1434–1441, 2011.
[75] C. P. Briscoe, M. Tadayyon, J. L. Andrews et al., “The orphan [92] H. Sun, Y. Hu, Z. Gu, R. T. Owens, Y. Q. Chen et al., “Omega-
G protein-coupled receptor GPR40 is activated by medium and 3 fatty acids induce apoptosis in human breast cancer cells and
long chain fatty acids,” Journal of Biological Chemistry, vol. 278, mouse mammary tissue through syndecan-1 inhibition of the
no. 13, pp. 11303–11311, 2003. MEK-Erk pathway,” Carcinogenesis, vol. 32, pp. 1518–1524, 2011.
[76] S. Hardy, G. G. St-Onge, E. Joly, Y. Langelier, and M. Prentki, [93] I. J. Edwards, I. M. Berquin, H. Sun et al., “Differential effects
“Oleate promotes the proliferation of breast cancer cells via of delivery of omega-3 fatty acids to human cancer cells by low-
the G protein-coupled receptor GPR40,” Journal of Biological density lipoproteins versus albumin,” Clinical Cancer Research,
Chemistry, vol. 280, no. 14, pp. 13285–13291, 2005. vol. 10, no. 24, pp. 8275–8283, 2004.
10 BioMed Research International

[94] H. Sun, I. M. Berquin, and I. J. Edwards, “Omega-3 polyunsatu- [110] A. Ishikado, Y. Nishio, K. Morino et al., “Low concentration
rated fatty acids regulate syndecan-1 expression in human breast of 4-hydroxy hexenal increases heme oxygenase-1 expression
cancer cells,” Cancer Research, vol. 65, no. 10, pp. 4442–4447, through activation of Nrf2 and antioxidative activity in vascular
2005. endothelial cells,” Biochemical and Biophysical Research Com-
[95] H. Sun, Y. Hu, Z. Gu et al., “Endogenous synthesis of n-3 munications, vol. 402, no. 1, pp. 99–104, 2010.
polyunsaturated fatty acids in Fat-1 mice is associated with [111] H. Wang, T. O. Khor, C. L. L. Saw et al., “Role of Nrf2 in
increased mammary gland and liver Syndecan-1,” PLoS ONE, suppressing LPS-induced inflammation in mouse peritoneal
vol. 6, no. 5, Article ID e20502, 2011. macrophages by polyunsaturated fatty acids docosahexaenoic
[96] Y. H. Teng, R. S. Aquino, and P. W. Park, “Molecular functions acid and eicosapentaenoic acid,” Molecular Pharmaceutics, vol.
of syndecan-1 in disease,” Matrix Biology, vol. 31, pp. 3–16, 2012. 7, no. 6, pp. 2185–2193, 2010.
[97] T. Ishikawa and R. H. Kramer, “Sdc1 negatively modulates car- [112] M. J. Magbanua, R. Roy, E. V. Sosa, V. Weinberg, S. Federman et
cinoma cell motility and invasion,” Experimental Cell Research, al., “Gene expression and biological pathways in tissue of men
vol. 316, no. 6, pp. 951–965, 2010. with prostate cancer in a randomized clinical trial of lycopene
and fish oil supplementation,” PLoS One, vol. 6, no. 9, Article ID
[98] J. Kiviniemi, M. Kallajoki, I. Kujala et al., “Altered expression of e24004, 2011.
syndecan-1 in prostate cancer,” APMIS, vol. 112, no. 2, pp. 89–97,
[113] B. Minke, “TRP channels and Ca2+ signaling,” Cell Calcium, vol.
2004.
40, no. 3, pp. 261–275, 2006.
[99] D. Chen, B. Adenekan, L. Chen et al., “Syndecan-1 expression [114] J. A. Matta, R. L. Miyares, and G. P. Ahern, “TRPV1 is a
in locally invasive and metastatic prostate cancer,” Urology, vol. novel target for omega-3 polyunsaturated fatty acids,” Journal
63, no. 2, pp. 402–407, 2004. of Physiology, vol. 578, no. 2, pp. 397–411, 2007.
[100] T. Zellweger, C. Ninck, M. Mirlacher et al., “Tissue microarray [115] M. Parnas, M. Peters, and B. Minke, “Linoleic acid inhibits TRP
analysis reveals prognostic significance of syndecan-I expres- channels with intrinsic voltage sensitivity: implications on the
sion in prostate cancer,” Prostate, vol. 55, no. 1, pp. 20–29, 2003. mechanism of linoleic acid action,” Channels, vol. 3, no. 3, pp.
[101] I. J. Edwards, H. Sun, Y. Hu et al., “In vivo and in vitro regulation 164–166, 2009.
of syndecan 1 in prostate cells by n-3 polyunsaturated fatty [116] M. Leonelli, M. F. Graciano, and L. R. Britto, “TRP channels,
acids,” Journal of Biological Chemistry, vol. 283, no. 26, pp. omega-3 fatty acids, and oxidative stress in neurodegeneration:
18441–18449, 2008. from the cell membrane to intracellular cross-links,” Brazilian
[102] P. A. Corsetto, G. Montorfano, S. Zava, I. E. Jovenitti, A. Journal of Medical and Biological Research, vol. 44, pp. 1088–
Cremona, and A. M. Rizzo, “Effects of n-3 PUFAs on breast 1096, 2011.
cancer cells through their incorporation in plasma membrane,” [117] S. Ye, L. Tan, J. Ma, Q. Shi, and J. Li, “Polyunsaturated docosa-
Lipids in Health and Disease, vol. 10, article 73, 2011. hexaenoic acid suppresses oxidative stress induced endothelial
[103] J. B. Ewaschuk, M. Newell, and C. J. Field, “Docosahexanoic acid cell calcium influx by altering lipid composition in membrane
improves chemotherapy efficacy by inducing CD95 transloca- caveolar rafts,” Prostaglandins Leukotrienes and Essential Fatty
tion to lipid rafts in ER(-) breast cancer cells,” Lipids, vol. 47, pp. Acids, vol. 83, no. 1, pp. 37–43, 2010.
1019–1030, 2012. [118] M. Flourakis and N. Prevarskaya, “Insights into Ca2+ homeosta-
[104] J. Y. Lee, L. Zhao, and D. H. Hwang, “Modulation of pat- sis of advanced prostate cancer cells,” Biochimica et Biophysica
tern recognition receptor-mediated inflammation and risk of Acta, vol. 1793, no. 6, pp. 1105–1109, 2009.
chronic diseases by dietary fatty acids,” Nutrition Reviews, vol. [119] Y. Sun, S. Selvaraj, A. Varma, S. Derry, A. E. Sahmoun et al.,
68, no. 1, pp. 38–61, 2010. “Increase in serum Ca2+ /Mg2+ ratio promotes proliferation of
[105] Y. Mitsuishi, H. Motohashi, and M. Yamamoto, “The prostate cancer cells by activating TRPM7 channels,” Journal of
Keap1-Nrf2 system in cancers: stress response and anabolic Biological Chemistry, vol. 288, pp. 255–263, 2013.
metabolism,” Frontiers in Oncology, vol. 2, article 200, 2012.
[106] H. Wang, T. O. Khor, C. L. L. Saw et al., “Role of Nrf2 in
suppressing LPS-induced inflammation in mouse peritoneal
macrophages by polyunsaturated fatty acids docosahexaenoic
acid and eicosapentaenoic acid,” Molecular Pharmaceutics, vol.
7, no. 6, pp. 2185–2193, 2010.
[107] P. Zhang, A. Singh, S. Yegnasubramanian et al., “Loss of
kelch-like ECH-associated protein 1 function in prostate cancer
cells causes chemoresistance and radioresistance and promotes
tumor growth,” Molecular Cancer Therapeutics, vol. 9, no. 2, pp.
336–346, 2010.
[108] A. Singh, S. Boldin-Adamsky, R. K. Thimmulappa et al., “RNAi-
mediated silencing of nuclear factor erythroid-2-related fac-
tor 2 gene expression in non-small cell lung cancer inhibits
tumor growth and increases efficacy of chemotherapy,” Cancer
Research, vol. 68, no. 19, pp. 7975–7984, 2008.
[109] L. Gao, J. Wang, K. R. Sekhar et al., “Novel n-3 fatty acid oxi-
dation products activate Nrf2 by destabilizing the association
between Keap1 and Cullin3,” Journal of Biological Chemistry, vol.
282, no. 4, pp. 2529–2537, 2007.
International Journal of

Peptides

Advances in
BioMed
Research International
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014
Stem Cells
International
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014
Virolog y
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014
International Journal of
Genomics
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014

Journal of
Nucleic Acids

Zoology
International Journal of

Hindawi Publishing Corporation Hindawi Publishing Corporation


http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014

Submit your manuscripts at


http://www.hindawi.com

Journal of The Scientific


Signal Transduction
Hindawi Publishing Corporation
World Journal
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014

Genetics Anatomy International Journal of Biochemistry Advances in


Research International
Hindawi Publishing Corporation
Research International
Hindawi Publishing Corporation
Microbiology
Hindawi Publishing Corporation
Research International
Hindawi Publishing Corporation
Bioinformatics
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014

Enzyme International Journal of Molecular Biology Journal of


Archaea
Hindawi Publishing Corporation
Research
Hindawi Publishing Corporation
Evolutionary Biology
Hindawi Publishing Corporation
International
Hindawi Publishing Corporation
Marine Biology
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014

You might also like