ArticlePDF AvailableLiterature Review

Chemo-preventive and therapeutic effect of the dietary flavonoid kaempferol: A comprehensive review

Authors:
REVIEW
Chemopreventive and therapeutic effect of the dietary
flavonoid kaempferol: A comprehensive review
Muhammad Imran
1
|Abdur Rauf
2
|Zafar Ali Shah
2
|Farhan Saeed
3
|Ali Imran
3
|
Muhammad Umair Arshad
3
|Bashir Ahmad
4
|Sami Bawazeer
5
|Muhammad Atif
6
|
Dennis G. Peters
7
|Mohammad S. Mubarak
8
1
University Institute of Diet & Nutritional
Sciences, Faculty of Allied and Health
Sciences, The University of LahorePakistan
2
Department of Chemistry, University of
Swabi Anbar, Swabi, Pakistan
3
Faculty of Home and Food Sciences,
Government College University, Faisalabad,
Pakistan
4
Center of Biotechnology and Microbiology,
University of Peshawar, Peshawar, Pakistan
5
Department of EMS. Paramedic, College of
Public Health and Health Informatics, Umm
AlQura University, Makkah, Saudi Arabia
6
Department of Clinical Laboratory Sciences,
College of Applied Medical Sciences, Jouf
University, Sakaka, Saudi Arabia
7
Department of Chemistry, Indiana University,
Bloomington, Indiana, USA
8
Department of Chemistry, The University of
Jordan, Amman, Jordan
Correspondence
Muhammad Imran, University Institute of Diet
and Nutritional Sciences, Faculty of Allied and
Health Sciences, The University of Lahore,
Lahore, Pakistan.
Email: mic_1661@yahoo.com
Abdur Rauf, Department of Chemistry,
University of Swabi Anbar23430, KPK,
Pakistan.
Email: mashaljcs@yahoo.com
Mohammad S. Mubarak, Department of
Chemistry, The University of Jordan, Amman
11942, Jordan.
Email: mmubarak@ju.edu.jo
Kaempferol, a natural flavonoid present in several plants, possesses a wide range of
therapeutic properties such as antioxidant, anticancer, and antiinflammatory. It has
a significant role in reducing cancer and can act as a therapeutic agent in the treat-
ment of diseases and ailments such as diabetes, obesity, cardiovascular diseases, oxi-
dative stress, asthma, and microbial contamination disorders. Kaempferol acts through
different mechanisms: It induces apoptosis (HeLa cervical cancer cells), decreases cell
viability (G2/M phase), downregulates phosphoinositide 3kinase (PI3K)/AKT (protein
kinase B) and human Tcell leukemia/lymphoma virusI (HTLVI) signaling pathways,
suppresses protein expression of epithelialmesenchymal transition (EMT)related
markers including Ncadherin, Ecadherin, Slug, and Snail, and metastasisrelated
markers such as matrix metallopeptidase 2 (MMP2). Accordingly, the aim of the
present review is to collect information pertaining to the effective role of kaempferol
against various degenerative disorders, summarize the antioxidant, antiinflammatory,
anticancer, antidiabetic, and antiaging effects of kaempferol and to review the prog-
ress of recent research and available data on kaempferol as a protective and chemo-
therapeutic agent against several ailments.
KEYWORDS
anticancer, antidiabetic, cardioprotective, kaempferol, mechanisms of action, oxidative stress
1|INTRODUCTION
Health benefits related to consumption of fruits and vegetables
attracted the attention of researchers and investigators owing to the
presence of bioactive compounds and to their effect on human health
(Ghosh, Dey, & Saha, 2014). These compounds exert anticancer, anti-
diabetic, antiobesity, cardioprotective, antiinflammatory, antiallergic,
and antiplatelet activities (Jose, Sudhakaran, Kumar, Jayaraman, &
Variyar, 2014). Flavonoids are hydroxylated phenolic substances hav-
ing a benzoγpyrone structure. They are widely present in plants in
response to microbial infection and are synthesized by a
phenylpropanoid pathway (Mahomoodally, GuribFakim, & Subratty,
2005). Health endorsing perspectives of flavonoids have been corre-
lated with their intake. According to estimates, the intake in humans
ranges from 20 to 1,000 mg/day, depending upon the consumption
pattern of the population (Scalbert & Williamson, 2000; Wang et al.,
Received: 27 July 2018 Revised: 24 September 2018 Accepted: 16 October 2018
DOI: 10.1002/ptr.6227
Phytotherapy Research. 2018;113. © 2018 John Wiley & Sons, Ltd.wileyonlinelibrary.com/journal/ptr 1
2009). As flavonoids like kaempferol are abundantly present in leafy
vegetables, apples, onions, broccoli, and berries, their consumption is
higher in all populations owing to the presence of this source in all
food guide models across the globe. Moreover, higher bioactivity
and therapeutic potential of flavonoids have been elucidated by their
higher heat stability and minor losses during cooking and frying oper-
ations. However, storage conditions alongside dietary interactions
may prove lethal for its bioavailability (Arts & Hollman, 2005; Moon,
Wang, & Morris, 2006).
Bioactivities of flavonoids depend mainly on their degree of
hydroxylation, structural class, other substitutions and conjugations,
and degree of polymerization (Heim, Tagliaferro, & Bobilya, 2002). In
addition, flavonoids exhibit antioxidant activity and neutralize the
effects of free radicals due to the presence of hydroxyl groups, and they
chelate metal ions (Kumar & Pandey, 2013). Recent epidemiological
findings suggest that higher dietary intake of flavonoid may be
inversely associated with risk of mortality (Godos, Castellano, Ray,
Grosso, & Galvano, 2018), diabetes (Liu et al., 2014), cardiovascular dis-
eases (Hooper et al., 2008), certain cancers (Grosso et al., 2017), and
depressive mood disorders (Chang et al., 2016; Godos et al., 2018).
Kaempferol (3,5,7trihydroxy2[4hydroxyphenyl]4H1benzopyran
4one, Figure 1) is a yellow bioactive flavonoid, which is present in
many edible plants such as tea, cabbage, broccoli, endive, kale, beans,
tomato, strawberries, leek, and grapes (CalderónMontaño, Burgos
Morón, PérezGuerrero, & LópezLázaro, 2011). As an antioxidant,
kaempferol counteracts production of superoxide ions and lowers the
formation of reactive oxygen and nitrogen species. It also scavenges
Fentongenerated hydroxyl radical, peroxynitrite, and hydroxyl radicals
(L. Wang et al., 2006). Furthermore, kaempferol suppresses the activity
of xanthine oxidase and enhances the activities of catalase, heme
oxygenase1, and superoxide dismutase (Heijnen, Haenen, Van Acker,
Van der Vijgh, & Bast, 2001; Klaunig & Kamendulis, 2004). Kaempferol
additionally exerts anticancer activity through several pathways such as
inhibition of angiogenesis and expression of vascular endothelial
growth factor (VEGF), induction of apoptosis, regulation of hypoxia
inducible factor 1alpha (HIF1α), induction of G2/M cell cycle arrest,
and caspase3dependent apoptosis (Huang et al., 2013; Kang et al.,
2010). Owing to the numerous uses of kaempferol as a chemothera-
peutic agent in the treatment of several ailments, this review focuses
on present knowledge about its effect in reducing the risk, preventing
and treating diseases (including cancer), along with its mechanisms of
action. Different databases such as MEDLINE (PubMed), Google
Scholar, Science Direct, and SciFinder have been employed to obtain
recent relevant references pertaining to the bioactivity and uses of
kaempferol as a chemotherapeutic agent by adapting the systematic
review methodology. Shown inTable 1 is a list of diseases treated with
or prevented by kaempferol, with mechanisms of action and a list of
related references.
2|HEALTH PERSPECTIVES
2.1 |Anticancer properties
Genomic DNA methylation plays a pivotal role in the progression and
development of bladder cancer. In this respect, kaempferol, as a che-
mopreventive agent, has been found to modulate DNA methylation,
to suppress the protein levels of DNA methyltransferases (DNMT3B)
and to induce 103 differential DNA methylation positions (dDMPs)
associated with genes 50 hypermethylated and 53 hypomethylated.
Additionally, it induces a premature degradation of DNMT3B by sup-
pressing the protein synthesis with cycloheximide (CHXv) (Qiu et al.,
2017). In a recent investigation, Kashafi and coworkers reported that
kaempferol (12100 μM) induces apoptosis and cell death in HeLa
cervical cancer cells in a dosedependent fashion. These researchers
also found that treatment with kaempferol leads to reduction in cell
viability and downregulation of PI3K/AKT and hTERT pathways
(Kashafi, Moradzadeh, Mohamadkhani, & Erfanian, 2017). Similarly,
in MCF7 breast cancer cell line, research findings revealed that
kaempferol, at a dose of 25 μM, inhibits the protein expression of
EMTrelated markers including Ncadherin, Slug, Ecadherin, and Snail
and suppresses metastasisrelated markers such as MMP2, 9, cathep-
sin B, and D (Lee, Choi, & Hwang, 2017). The antioncogenic role of
dietary flavonoid intake is further divulged by a metaanalysis in which
a systematic search in electronic databases published up to June 2016
(143) was performed. The outcomes showed that the flavonoids con-
sumption has significant effect in different kinds of cancer prevention
(Grosso et al., 2017).
In human umbilical vein endothelial cells (HUVECs), Lee et al.
(2016) showed that kaempferol induces subG1 phase cell population,
activates the caspase signals such as caspase3, 8, and 9 and triggers
apoptosis in a dosedependent manner. Additionally, it stimulates
death receptor signals (death receptor 4 [DR4], Fas/CD95, and DR5)
by enhancing the expressions of phosphorylated ataxia telangiectasia
mutated (ATM) and phosphorylated p53 pathways (Lee et al., 2016).
In a similar fashion, research by Liao and colleagues revealed that
kaempferol displays anticancer activity against different human cancer
cell lines, such as the human stomach carcinoma (SGC7901), human
breast carcinoma (MCF7), human lung carcinoma (A549), and human
cervical carcinoma (Liao et al., 2016). On the other hand, Tu, Bai, Cai,
and Deng (2016) showed that orally administered kaempferol, dose
dependently suppresses the spread of human cervical cancer by induc-
ing cell apoptosis, intracellular free calcium elevation, and mitochon-
drial membrane potential disruption (Tu et al., 2016). In HCCC9810
and QBC939 cells, kaempferol suppressed proliferation, reduced
colony formation, and induced apoptosis. Additionally, it (a) decreased
Bcl2 expression, (b) increased Fas, Bax expressions, (c) enhanced
cleavedcaspase 3, cleavedcaspase 8, cleavedcaspase 9, and
cleavedPARP, (d) downregulated levels of phosphorylated TIMP2,
AKT, and MMP2, and (e) inhibited the number and volume of metasta-
sis foci in lung metastasis model (Qin, Cui, Yang, & Tong, 2016).
Similarly, research findings showed that kaempferol markedly sup-
presses 17βestradiol (E2), induces apoptosis, causes cell cycle arrest,
upregulates the expressions of bax and p21, downregulates the
expressions of cyclin E, cyclin D1, and cathepsin D. It additionally
FIGURE 1 Structure of kaempferol
2IMRAN ET AL.
TABLE 1 Health perspectives of kaempferol
Disorders Mechanisms References
Anticancer Modulation of DNA methylation, induced 103 differential DNA methylation positions
(dDMPs).
Suppression of protein levels of DNA methyltransferases (DNMT3B)
(Qiu et al., 2017).
Reduces cell viability, and downregulates PI3K/AKT and hTERT pathways. (Kashafi et al., 2017).
Inhibition of Ncadherin, Ecadherin, Slug, Snail, and MMP2, 9, and cathepsin B, D. (Lee et al., 2017).
Induction of subG1 phase cell population.
Activation of caspase signals such as caspase3, 8, and 9
(Lee et al., 2016).
Reduces levels of Bcl2, enhances Fas and Bax expressions, and increases cleaved
caspase 3, cleavedcaspase 8, cleavedcaspase 9, and cleavedPARP.
Downregulation of phosphorylated TIMP2, AKT, and MMP2 levels.
(Qin et al., 2016).
Induction of cell apoptotic cell death, intracellular free calcium elevation, and
mitochondrial membrane potential disruption.
Tu et al., 2016
Oxidative Stress Suppresses the activity of ASK1/MAPK signaling pathways (JNK1/2 and p38). (Feng et al., 2017;
Choung et al., 2017).
Enhances the concentrations of superoxide dismutase, catalase, glutathione peroxidase,
and glutathioneStransferase.
Increases plasma insulin level.
(AlNumair, Veeramani,
Alsaif, & Chandramohan,
2015).
Lowers aspartate aminotransferase, alanine aminotransferase, malondialdehyde (MDA).
Decreases activity of hepatic microsomal enzyme cytochrome 2E1 (CYP2E1) expression.
(Wang et al., 2015)
Enhancement of Bid, Bcl2, MAPK, and AIF. (Yang et al., 2014)
Increases mRNA and protein expression of Nrf2regulated genes. (Saw et al., 2014).
Cardiovascular
Role
Suppresses advanced glycation end products (AGEs)receptor.
Reduces levels of IL6, TNFα, and NFκB.
Activates ERK1/2 and inhibits active cJNK and p38 proteins.
Lowers expressions of Caspase3 and Bax, TUNEL positive cells.
(Suchal et al., 2017).
Reduces levels of proapoptotic proteins (Caspase3 and Bax), TUNEL positive cells. (Suchal, et al., 2016)
Reduces levels of cytoplasm cytochrome C, cleaved caspase3, lactate dehydrogenase,
and creatine kinase
Lowers tumor necrosis factoralpha (TNFα), and malondialdehyde
Zhou et al., 2015
Decreases expression of nicotinamide adenine dinucleotide phosphateoxidase 4 (Nox4),
OPN, NFκB, inhibitor of NFκB alpha phosphorylation (PIκBα), IL6, and alphavbeta3
(αvβ3) integrin
(Xiao et al., 2016;
Yang et al., 2014).
Antidiabetic Inhibits cell proliferation, migration, migration distance, and sprouting.
Downregulates the expression of PI3K
Inhibits the activation of Src, Erk1/2, and Akt1.
(Xu et al., 2017)
Lowers p62 expression, up regulated the AMPK, and downregulates the mTOR
phosphorylation.
(Varshney et al., 2017;
Zeng et al., 2015)
Decreases SREBP1c expression.
Lowers PPARγ, and serum HbA1c levels.
Zeng et al., 2015)
Suppresses the phosphorylation of insulin receptor substrate1 (IRS1), IkB kinase β
(IKKβ), and IkB kinase α(IKKα).
(Luo et al., 2015;
Peng et al., 2016)
Antiinflammatory Inhibits the release of IL1β, TNFα,IL18, and IL6. (Tang et al., 2015)
Decreases expressions of TSLP, inhibits the upregulation of tolllike receptor 4 (TLR4),
phosphorylation level of IκBα.
(Nam et al., 2017).
Suppresses myeloid differentiation factor 88 (MyD88), and NFκB p65 DNA binding
activity.
(Zhang et al., 2017).
Antiaging Improves motor coordination, and enhanced striatal dopamine. (Li & Pu, 2011).
Inhibits fibrilogenesis and secondary structural transformation of the peptide. (Sharoar et al., 2012)
Antiallergic Inhibits COX2mediated production of prostaglandin D2 and prostaglandin F2α.
Lowers αSMA expression, blocked BSA inhalationinduced epithelial cell excrescence
and smooth muscle hypertrophy.
Dampens the antigenchallenged activation of Sykphospholipase Cγ(PLCγ) pathway.
(Shin et al., 2015)
Inhibits enhancement of TNFαprotein levels and Th2 cytokines (IL4, IL5, and IL13).
Suppresses phosphorylation Akt and eosinophilia.
(Chung et al., 2015)
Inhibits Mucin gene expression and represses PQ phosphorylation of ERKs and cJNK. (Podder et al., 2014).
Inhibits tunicamycininduced ER stress and attenuates the induction of XBP1 and
IRE1α.
(Park et al., 2015)
Antiplatelet
aggregation
Inhibits enzymatic activities of thrombin, FXaas, and fibrin polymer formation.
Attenuates the phosphorylation of ERK1/2, p38, cJNK1/2, and phosphoinositide 3
kinase (PI3K)/PKB (AKT).
(Choi et al., 2015)
Bone disorders Induces apoptotic cell death and decreases viability of mitosisassociated nuclear antigen
(Ki67).
Lowers levels of MMP3 and MMP13 enzymes.
(Zhu et al., 2017)
Suppresses IL1βstimulated and RANKLmediated osteoclast differentiation
Inhibits RANKLmediated phosphorylation of ERK 1/2, JNK MAP kinases, p38, and
expressions of NFATc1 and cFos
(Lee et al., 2014;
Nepal et al., 2013).
Suppresses the proliferation and migration of VSMC
Activates BMP signaling pathway, downregulates DOCK4, 5, and 7, and induces miR21
expression.
(Kim et al., 2015)
(Kim et al., 2016)
(Continues)
IMRAN ET AL.3
caused reduction of the phosphorylation of AKT, IRS1, ERK, and
MEK1/2 in MCF7 breast cancer cells of in vivo xenografted mouse
model (Jo, Park, Choi, Jeon, & Kim, 2015; Kim, Hwang, & Choi,
2016). In addition, Lee and Kim (2016) reported that kaempferol
exhibits anticancer activity against different human pancreatic cancer
cell lines such as Miapaca2, Panc1, and SNU213 in a dose
dependent fashion. Moreover, it lowered viability, induced apoptosis,
inhibited migratory activity, and mediated by inhibition of ERK1/2,
EGFRrelated Src, and AKT pathways (Lee & Kim, 2016). In regulatory
T cells (Tregs), kaempferol significantly enhanced the inhibitory effect
of proliferation, increased the FOXP3 expression level, protected from
the pathological symptoms of collageninduced arthritis, and also
decreased the PIM1mediated FOXP3 phosphorylation at S422 in
rat model (Lin et al., 2015).
In human leukemia HL60 cells, kaempferol dosedependently
exhibited multiple mechanisms such as induction of DNA damage,
enhancement DNA condensation (Comet tail), reduction of
phosphateataxiatelangiectasia, protein expression associated with
DNA repair system, including phosphateataxiatelangiectasia mutated
(pATM), DNAdependent serine/threonine protein kinase (DNAPK),
Rad3related (pATR), O(6)methylguanineDNA methyltransferase
(MGMT), 1433 proteins sigma (1433σ), p53, and MDC1 protein
expressions, as well as enhanced protein expression of pH2AX and
pp53 (Wu et al., 2015). Similarly, findings of Azevedo et al. (2015)
highlighted the protective role of kaempferol against the proliferation
of MCF7 cells through inhibition of HdeoxyDglucose (3 HDG)
uptake, lowering GLUT1 mRNA levels, increasing extracellular lactate
levels, and suppressing MCT1mediated lactate cellular uptake
(Azevedo et al., 2015). In human fibrosarcoma HT1080 cells,
kaempferol at a concentration of 30 μM showed significant reduction
in matrix metalloproteinase9 (MMP)9 secretion, JNK phosphoryla-
tion, and IκBαphosphorylation and inhibited phorbol12myristate
13acetate (PMA)induced MMP9 expression by blocking activation
of AP1 and NFκB (Choi et al., 2015).
In a similar fashion, treatment with kaempferol inhibited prolifera-
tion of gastric cancer cell lines MKN28 and SGC7901 through multiple
mechanisms; it (a) induced apoptotic cell death, (b) exhibited cell cycle
arrest, (c) suppressed tumor growth, (d) decreased cyclin B1, Cdk1, and
Cdc25C expressions, (e) lowered Bcl2 and enhanced Bax expressions,
(f) upregulated cleaved caspase3 and 9, (g) promoted Poly (ADP
ribose) polymerase (PARP) cleavage, and (h) decreased pAkt, pERK
and cyclooxygenase2 (COX2) expression levels (Song et al., 2015).
Furthermore, kaempferol3Orhamnosidedose, a derivative of
kaempferol, suppressed cell proliferation rate, induced caspase
cascade pathway, upregulated the expression of caspase3, caspase
8, caspase9, and PARP proteins in LNCaP human prostate cancer cell
lines in a concentrationdependent manner (Halimah et al., 2015).
Treatment of HT29 rat cells with different doses of kaempferol
(060 μmol/L) resulted in induction of apoptosis, an increase in chro-
matin condensation, and fragmentation of DNA in a dosedependent
manner. It also increased the levels of cleaved caspase3 and
caspase7 and cleaved PARP. Furthermore, it caused (a) enhancement
in mitochondrial membrane permeability and cytosolic cytochrome c
concentrations, (b) reduction in BclxL proteins expression, (c) increase
of the proapoptotic protein Bik, (d) reduction of Akt activation and
Akt activity, (e) enhancement of mitochondrial Bad, (f) increase of
membranebound FAS ligand, (g) decrease of uncleaved caspase8
activity, and (h) enhancement of caspase8 activity (Lee et al., 2014).
In a similar fashion, Cho and Park (2013) conducted a study on the
proliferation of HT29 cancer cells treated with different doses of
kaempferol (060 μmol/L). These researchers found that kaempferol
significantly decreases viable cell numbers, [
3
H] thymidine incorpora-
tion into DNA, causes G1 and G2/M cell cycle arrest, inhibits protein
expressions (cyclin A, cyclin E, cyclins D1, CDK2, CDK4), inhibits
CDK2 and CDK4 activities, and reduces phosphorylation of retino-
blastoma protein. Moreover, it lowers expressions of Cdc2, Cdc25
C`, and cyclin B1 proteins (Cho & Park, 2013).
A paper published by Yao et al. (2014) explored the inhibitory role
of kaempferol against human squamous cell carcinoma (SCC) and
SUVtreated mouse skin. These researchers found that kaempferol
inhibits mitogen and stressactivated protein kinase (MSK) and 90
kDa ribosomal S6 kinase (RSK) proteins, lowers SUVinduced phos-
phorylation of histone H3 and cFos, and attenuates SUVinduced
phosphorylation of cAMPresponsive element binding protein (CREB)
and histone H3 (Yao et al., 2014). On the other hand, researchers
found that kaempferol exhibits a cytotoxic effect on the proliferation
of bladder cancer cells by inducing apoptosis, causing cell cycle arrest,
inhibiting tumor growth, lowering levels of growthrelated markers,
and enhancing the expression of apoptosis markers. Moreover, it sup-
presses invasion and metastasis and downregulates the cMet/p38
signaling pathway (Dang et al., 2015; Nepal et al., 2013). Kaempferol
additionally reduces the viability, tumor growth, and tumor size in dif-
ferent human cancer cell lines including pharynx (FaDu cell line), met-
astatic lymph node (PCI15B cell line), oral cavity carcinoma (PCI13
cell line), and explanted FaDu cells (Swanson, Choi, Helton, Gairola,
& Valentino, 2014). Other studies verified the preventive role of
kaempferol against renal cell carcinoma (RCC) lines (786O and 769
P cells) through multiple mechanisms such as inhibition of cell growth,
induction of apoptotic cell death and cell cycle arrest, upregulation of
p21 and downregulation of cyclin B1 expressions, inhibition of
TABLE 1 (Continued)
Disorders Mechanisms References
Increases expression of the osteoblastactivated factors RUNX2, BMP2, osterix,
collagen I, and SQSTM1/p62
Antiobesity Reduces expressions of lipin1, FASN, LPAATθ(lysophosphatidic acid acyltransferase),
SREBP1C (fatty acid synthetic proteins), and DGAT1 (triglyceride synthetic enzymes).
Blocks the mammalian target of mTOR, and phosphorylation of AKT (protein kinase B).
Attenuates C/EBPαand peroxisome proliferatoractivated receptor γ(PPARγ).
(Lee et al., 2015).
Lowers the expression of adipogenic transcription factors (Pparγ, Cebpβ, Srebp1, Rxrβ,
Lxrβ, and Rorα)
(Park et al., 2012).
4IMRAN ET AL.
activation of EGFR/p38 signaling pathways, and activation of PARP
cleavages (Song et al., 2014). Flavonoids such as kaempferol are more
promising in cancer inhibition then the flavanols owing to their struc-
tural diversity. In this context, the number of OH on the B ring influ-
ences the binding and inhibition (Debashis et al., 2017). Likewise,
observations made by Das, Majumder, and Saha (2017) concluded that
the DNA binding ability of flavonols has been attributed to their struc-
ture. Moreover, Pal, Dey, and Saha (2014) showed another mechanism
by which flavonoids may be helpful in the prevention of cancer. They
examined the catalase inhibitory activity of green tea and suggested
that its anticancer property is mainly defined by ROS accumulation
due to catalase inhibition as depicted in Figure 2.
2.2 |Cardiovascular effect
A paper published by Suchal and coworkers (Suchal et al., 2017) indi-
cated that kaempferol is effective against myocardial ischemia
reperfusion (IR) injury in diabetic male albino Wistar rats. It markedly
decreases hyperglycemia, suppresses advanced glycation end prod-
ucts (AGEs)receptor for advanced glycation end products (RAGE) axis
activation, maintains hemodynamic function, preserves morphological
alterations, and normalizes oxidative stress. It additionally lowers IL
6, TNFα, and NFκB levels, activates ERK1/2, and inhibits cJNK
and p38 proteins. Likewise, kaempferol attenuated apoptosis by
reducing the expression of proapoptotic proteins including caspase
3 and Bax and by increasing the level of antiapoptotic protein Bcl2.
Additionally, it enhanced the level of antiapoptotic protein (Bcl2). In
an earlier publication, Suchal et al. (2016) found that kaempferol (a)
lowers the expressions of IL6, TNFα, and NFκB, (b) suppresses p38
and JNK proteins, (c) activates ERK1/ERK2, (d) induces apoptosis, (e)
reduces levels of proapoptotic proteins (caspase3 and Bax), TUNEL
positive cells, and (f) increases the level of antiapoptotic proteins
(Bcl2) in the IR model of myocardial injury (Suchal et al., 2016).
Hooper and coworkers conducted a metaanalysis review to correlate
the cardiovascular protective role of flavonoids. They included data
from MEDLINE, EMBASE, and Cochrane databases that were further
processed by inclusion and exclusion validity and finally carried out
the metaanalysis. They included 133 clinical trials and confirmed that
the dose dependent cardioprotective effect was observed for differ-
ent flavonoids (Hooper et al., 2008).
Regulation of osteopontin (OPN)related signaling pathways is
linked with aldosterone hormone by promoting activation of NFκB
in primary human umbilical vein endothelial cells (HUVECs). On the
other hand, kaempferol was found to lower production of ROS,
TNFα, OPN, NFκB, inhibitor of NFκB alpha phosphorylation (P
IκBα), IL6, and reduced nicotinamide adenine dinucleotide
phosphateoxidase 4 (Nox4), αvβ3 integrin, and PIκBαexpressions
(Xiao, Lu, Liu, & Luo, 2016). In a similar fashion, Zhou et al. (2015)
showed that kaempferol prevents myocardial I/R injury in rats by
decreasing the levels of TUNELpositive cell rate and by reducing
myocardial infarct size. It also reduces levels of cytoplasm cytochrome
C, cleaved caspase3, lactate dehydrogenase, creatine kinase, TNFα,
and malondialdehyde and improves the left ventricular developed
pressure and its maximum up/down rate (±dp/dt max). Moreover,
kaempferol increased the concentrations of glutathione/glutathione
disulfide, superoxide dismutase, phosphoGSK3β(PGSK3β), and
total glycogen synthase kinase3β(Zhou et al., 2015).
Research findings revealed that kaempferol prevents
anoxia/reoxygenation (A/R)induced injury of cardiomyocytes by
increasing cell viability, lowering LDH release, reducing A/Rinduced
ROS generation, loss of Δψm, and release of cytochrome c from mito-
chondria into cytosol. Furthermore, it upregulated human silent infor-
mation regulator Type 1 expression, suppressed the A/Rstimulated
mPTP opening, activated caspase3, and enhanced the expression of
Bcl2 (Guo et al., 2015). Similarly, it repressed the vascular smooth
muscle cell (VSMC) proliferation and migration and modulated the
microRNA expression levels and BMP4 signaling pathway. It addition-
ally induced apoptotic cell death, activated the BMP signaling path-
way, downregulated DOCK4, 5, and 7, induced miR21 expression,
and antagonized PDGFmediated promigratory effect (Kim et al.,
2015). Moreover, intraperitoneal administration of streptozotocin
(STZ; 40 mg/kg BW) induced diabetes in adult male albino rats,
whereas kaempferol (100 mg/kg BW) exhibited significant reduction
in activities of total ATPases, Na(+)/K(+)ATPase, Ca(2+)ATPase, and
Mg(2+)ATPase in erythrocytes and tissues (AlNumair et al., 2015).
FIGURE 2 Anticancer mechanism illustration of kaempferol showing its effects on the MAPK Pathway, cell cycle, and angiogenesis. [Colour
figure can be viewed at wileyonlinelibrary.com]
IMRAN ET AL.5
In myocardial infarction (MI) of male Wistar rats, oral intake of
kaempferol (5, 10, and 20 mg/kg/day, i.p.) caused substantial enhance-
ment in arterial pressure, reduction in left ventricular pressure, and left
ventricular enddiastolic pressure and increased the concentrations of
catalase, superoxide dismutase, and glutathione. Reduction in levels of
malondialdehyde, serum IL6, TNFαlevels, cardiac injury markers
(creatine kinaseMB and lactate dehydrogenase), and Bax/Bcl2 ratio
were reported after treatment with kaempferol (Suchal et al., 2016a).
In a similar fashion, a group of researchers investigated the effective
role of kaempferol on the relaxation of porcine coronary artery
smooth muscle cells. These researchers found that kaempferol
increases relaxations, which are endothelium derived, produced by
exogenous NO or due to endotheliumdependent hyperpolarization,
possibly by activation of KCa 1.1 channels (Xu, Leung, Leung, &
Man, 2015). On the other hand, Tang and colleagues investigated
the effect of kaempferol (12.5 and 25 μg/mL) on lipopolysaccharide
(LPS) plus ATPinduced cardiac fibroblasts and showed that it signifi-
cantly suppresses the release of IL1β, TNFα,IL6, and IL18 and
inhibits the activation of Akt and NFκB (Tang et al., 2015).
2.3 |Antidiabetic perspective
Diabetic retinopathy is one of the most prevalent microvascular com-
plications of diabetes. In human retinal endothelial cells (HRECs), glu-
cose at a concentration of 25 mM enhances the mRNA expression
levels of placenta growth factor (PGF) and VEGF, as well as the con-
centrations of secreted VEGF and PGF. Administration of kaempferol
at the rate of 525 μM, under highglucose conditions, markedly
inhibited cell proliferation, migration distance, and growth of HRECs.
Furthermore, kaempferol suppressed the activation of Src, Erk1/2,
and Akt1 and downregulated the expression of PI3K (Xu, Zhao, Peng,
Xie, & Liu, 2017). Varshney and coworkers recently studied the
cytoprotective effect of kaempferol against palmitic acidinduced pan-
creatic βcell death through modulation of autophagy via the
AMPK/mTOR signaling pathway. These researchers found that
kaempferol causes an increase in cell viability and apoptotic cell death
activities and improves the expressions of LC3 puncta and LC3II pro-
tein. These results suggest that kaempferol exerts a cytoprotective
role against lipotoxicity by activation of autophagy via the
AMPK/mTOR pathway (Varshney, Gupta, & Roy, 2017). Likewise,
the beneficial impact of dietary consumption of kaempferol has been
elucidated by the early metaanalysis carried out by Liu et al. (2014).
They carried out the metaanalysis to access relationship between dia-
betes onset and flavonoids consumption. They utilized a fixed effect
model to calculate the summary of risk and included the cohort stud-
ies from 2013. The outcomes significantly reflected the inverse asso-
ciation between flavonoids consumption and diabetes onset.
Research findings showed that treatment of diabetic rats with
kaempferol at doses of 50 and 150 mg/kg, dosedependently
improved blood lipids and insulin and suppressed the phosphorylation
of insulin receptor substrate1 (IRS1), IkB kinase β(IKKβ), and IkB
kinase α(IKKα) via the hepatic IKK/NFκB signaling pathway (Luo
et al., 2015; Peng, Zhang, Liao, & Gong, 2016). Furthermore, earlier
findings of Alkhalidy et al. (2015) indicated that kaempferol can cause
significant improvement in hyperglycemia, glucose tolerance, and
blood insulin levels in highfat fed obese mice. These results suggest
that kaempferol can act as an antidiabetic agent by improving periph-
eral insulin sensitivity and protecting against pancreatic βcell dysfunc-
tion (Alkhalidy et al., 2015). On the other hand, Zang and coworkers
found that treatment of C57BL/6J mice fed a highfat diet reverses
the effect of this food on parameters such as the adipose tissue, body
weight, triglyceride concentration, blood glucose level, peroxisome
proliferatoractivated receptor (PPARγ), sterol regulatory element
binding protein (SREBP1c) expression, and serum HbA1c (hemoglobin
A[1c]) levels (Zang, Zhang, Igarashi, & Yu, 2015). In INS1E cells and
human islets, Zhang and colleagues showed that kaempferol promotes
cell viability, suppresses apoptosis, and decreases caspase3 activity. It
additionally prevents downregulation of antiapoptotic proteins Bcl2
and Akt induced by lipotoxicity. Kaempferol also improved insulin
secretion and synthesis and increased pancreatic and duodenal
homeobox1 (PDX1) expression. Furthermore, it restored activation
of protein kinase A (PKA) and production of cyclic adenosine
monophosphate (cAMP) and CREB phosphorylation and its regulated
transcriptional activity in βcells via upregulation of the PDX1/
cAMP/PKA/CREB signaling pathway (Zhang et al., 2013).
2.4 |Oxidative stress
A recent publication by Feng and coworkers indicated that kaempferol
protects against cardiac hypertrophy by lowering cardiomyocyte areas
and interstitial fibrosis, along with improving cardiac functions and
decreasing apoptosis. Kaempferol exerts this protective role by
inhibiting the ASK1/MAPK signaling pathway and by regulating oxida-
tive stress (Feng, Cao, Zhang, & Wang, 2017). In a similar fashion, a
team of scientists investigated the hepatoprotective role of
kaempferol against alcoholic liver injury in mice. These researchers
found that kaempferol can significantly lower oxidative stress and lipid
peroxidation and can improve the antioxidative defense activity. In
addition, it considerably lowers both the expression level and activity
of hepatic CYP2E1 (Wang et al., 2015). In male albino Wistar rats,
alcoholinduced oxidative stress increases the level of γglutamyl
transferase (GGT) and lowers the concentrations of enzymatic antiox-
idants such as superoxide dismutase, catalase, and glutathione perox-
idase. However, research findings revealed that treatment with
kaempferol reverses these changes (Shakya, Manjini, Hoda, &
Rajagopalan, 2014). In STZinduced diabetic rats, kaempferol at a dose
of 100 mg/kg BW lowered plasma glucose, thiobarbituric acid reactive
substances, lipid hydroperoxides, conjugated dienes, and levels of vita-
mins E and C. It additionally increased glutathioneStransferase and
insulin plasma level (AlNumair et al., 2015).
Saw and coworkers studied the antioxidative stress role of
kaempferol, quercetin, and pterostilbene, alone and in combination,
in addition to the involvement of the Nrf2ARE signaling pathway.
These researchers showed that each of these compounds exhibited
remarkable freeradical scavenging activity in the DPPH assay,
attenuated intracellular ROS levels, and had synergistic effects when
used in combination. Furthermore, these compounds induced antioxi-
dant response element (ARE) and improved the mRNA and protein
6IMRAN ET AL.
expression of Nrf2regulated genes (Saw et al., 2014). In a similar fash-
ion, Yang and colleagues investigated the therapeutic role of
kaempferol against glutamatetreated hippocampal neuronal cells
(HT22). These researchers found that supplementation with
kaempferol at the rate of 25 μM shows substantial enhancement in cell
viability, regulation of expression levels of proteins such as Bid, Bcl2,
mitogenactivated protein kinase (MAPK), and apoptosisinducing
factor (Yang, Kim, Jun, & Song, 2014).
In male Wistar rats, research by Haidari, Keshavarz, Shahi,
Mahboob, and Rashidi (2011) showed that kaempferol (5 mg/kg)
reduces the levels of serum uric acid and inhibits the activity of xan-
thine oxidoreductase in a dosedependent fashion (Haidari et al.,
2011). On the other hand, Gao and coworkers examined the antioxi-
dative and antiapoptotic properties, along with the chemoprotective
mechanism of kaempferol. In House Ear InstituteOrgan of Corti 1
(HEIOC1) cells, these workers showed that kaempferol protects cells
against cisplatininduced apoptosis in a dosedependent manner.
Additionally, kaempferolinduced HO1 expression protected against
cell death though the cJun Nterminal kinase (JNK) pathway and
through Nrf2 translocation. Furthermore, kaempferol enhanced the
concentration of glutathione and the expression of glutamatecysteine
ligase catalytic (GCLC) subunit (Gao et al., 2010). Similarly, kaempferol
protected beta cells from 2deoxyDribose (dRib)induced oxidative
damage and suppressed dRibinduced intracellular ROS, apoptosis,
and lipid peroxidation (Lee et al., 2010). Likewise, a group of
researchers (Lagoa et al., 2009; Qu et al., 2009) reported that supple-
mentation of kaempferol provides protection against 3nitropropionic
acid (NPA)induced neurodegeneration in Wistar rats. In addition,
kaempferol reduced motor deficit, delayed mortality, and increased
levels of glutathione and protein nitrotyrosines. Furthermore, a study
by Lopez and colleagues highlighted the preventive role of kaempferol
against nitrosativeoxidative stress through enhancing apoptotic cell
death and PARP degradation and caspase9 activity (LópezSánchez
et al., 2007).
2.5 |Antiinflammatory properties
Kaempferol exhibits antiinflammatory activity and thus has a thera-
peutic potential for the treatment of inflammatory diseases. In human
airway epithelial BEAS2B cells and eosinophils, kaempferol markedly
inhibited LPSinduced eotaxin1 protein expression, attenuated
TNFαinduced expression of epithelial intracellular cell adhesion
molecule1 and eosinophil integrin β2, and diminished TNFαinduced
airway inflammation by attenuating monocyte chemoattractant
protein1 transcription, possibly by disturbing NFκB signaling (Gong,
Shin, Han, Kim, & Kang, 2012). In a similar fashion, Tang and
coworkers examined the antiinflammatory effect of kaempferol on
lipopolysaccharide (LPS) plus ATPinduced cardiac fibroblasts, along
with the underlying mechanisms of action. These researchers showed
that kaempferol considerably suppresses the release of TNFα,IL1β,
IL6, and IL18 and inhibits activation of NFκB and Akt in LPS plus
ATPinduced cardiac fibroblasts (Tang et al., 2015). In lipopolysaccha-
ride (LPS)treated RAW264.7 cells and peritoneal macrophages,
kaempferol exhibited inhibitory activity on the production of nitric
oxide (NO), COX2, and synthase (iNOS). It additionally reduced the
NFκBmediated enhancement of luciferase, activity, phosphorylation
of IκBα, and nuclear translocation of p50 and p65 (Jeong et al., 2014).
Recent in vivo and in vitro studies by Zhang et al. (2017) revealed
that kaempferol attenuates pulmonary edema, myeloperoxidase
activity, pulmonary capillary permeability, and numbers of inflamma-
tory cells. It additionally reduced the production of ROS and
malondialdehyde, enhanced superoxide dismutase activity, and
lowered the overproduction of IL1β, TNFα, and IL6. Furthermore,
it decreased H9N2 viral titer, and it inhibited the upregulation of toll
like receptor 4 (TLR4), phosphorylation level of IκBα, myeloid differen-
tiation factor 88 (MyD88), NFκB p65 DNA binding activity, NFκB
p65, and phosphorylation level of MAPKs. These results suggest that
kaempferol displays a protective effect on H9N2 virusinduced inflam-
mation by suppressing TLR4/MyD88mediated NFκB and MAPKs
pathways (Zhang et al., 2017). In addition, kaempferol has been found
to (a) attenuate IL32induced monocyte differentiation to product
macrophagelike cells, (b) decrease production and mRNA expression
of proinflammatory cytokines such as thymic stromal lymphopoietin
(TSLP), IL1β, TNFα, and IL8, (c) inhibit the IL32induced activation
of p38 and nuclear factorκB in THP1 cells, and (d) ameliorate the
lipopolysaccharideinduced production of the inflammatory mediators
TSLP, IL1β, TNFα,IL8, and nitric oxide of macrophagelike cells dif-
ferentiated by IL32 (Nam, Jeong, & Kim, 2017).
In LPStreated macrophages, sodium nitroprusside(SNP)treated
RAW264.7 cells and peritoneal macrophages, kaempferol inhibited
the release of prostaglandin E2 (PGE2) and nitric oxide (NO), neutral-
ized production of radicals, downregulated the cellular adhesion of
U937 cells to fibronectin (FN), and diminished mRNA expression levels
of inflammatory genes TNFα, encoding iNOS and COX2. In addition,
treatment with kaempferol caused a decrease in AP1(cFos and
cJun) and NFκB (p50 and p65) levels and inhibited Syk, Src, and
IRAK1 (S. H. Kim et al., 2015). Furthermore, Lin et al. (2015) explored
the antiinflammatory role of kaempferol against collageninduced
arthritis in experimental rats. Results showed that kaempferol
enhances the levels of FOXP3 in T cells and decreases PIM1mediated
FOXP3 phosphorylation at S422 (Lin et al., 2015). Shown in Figure 3 is
the antiinflammatory mechanism associated with kaempferol.
2.6 |Antiaging role
In the mouse model of Parkinson's disease, 1methyl4phenyl
1,2,3,6tetrahydropyridine (MPTP) induces the disease, whereas
kaempferol (a) enhanced the activity of glutathione peroxidase, cata-
lase, and superoxide dismutase, (b) reduced malondialdehyde levels,
(c) improved motor coordination, (d) elevated striatal dopamine and
its metabolite levels, and (e) prevented the loss of THpositive neurons
in the substantia nigra (Li & Pu, 2011). In pheochromocytoma cells
(PC12), kaempferol reversed amyloid beta peptide (Abeta)induced
impaired performance (Roth, Schaffner, & Hertel, 1999). Similarly,
kaempferol3Orhamnoside exhibits an antiamyloidogenic Aβ42
mediated cytotoxicity effect through several mechanisms such as sup-
pression of fibrilogenesis and secondary structural transformation of
the peptide (Sharoar et al., 2012).
IMRAN ET AL.7
2.7 |Antiallergic effect
Shin and coworkers employed an in vitro model of dinitrophenylated
bovine serum albumin (DNPBSA)sensitized rat basophilic leukemia
(RBL2H3) mast cells, and an in vivo model of BSAchallenged asth-
matic mice to explore the antiallergic effect of kaempferol. These
researchers found that kaempferol (a) markedly lowers antiαsmooth
muscle actin (αSMA) expression, (b) inhibits COX2mediated produc-
tion of prostaglandin D2 and prostaglandin F2α, (c) deters the antigen
induced mastcell activation of cytosolic phospholipase A2 response
to protein kinase Cμand extracellular signalregulated kinase (ERK),
(d) blocks bovine serum albumin (BSA) inhalationinduced epithelial
cell excrescence and smooth muscle hypertrophy, and (e) dampens
the antigenchallenged activation of the Sykphospholipase Cγ(PLCγ)
pathway (Shin et al., 2015). In similar fashion, Park et al. (2015) studied
the inhibitory role of kaempferol in the bronchial airway and lung of
BALB/c mice sensitized with ovalbumin (OVA). They showed that
orally administrated kaempferol is greater than or equal to 10 mg/kg
inhibits mucus secretion and goblet cell hyperplasia, whereas treat-
ment with 20 μM kaempferol dampened the TGFβand tunicamycin
promoted MUC5AC induction. Kaempferol additionally inhibited
tunicamycininduced ER stress and attenuated the induction of
XBP1 and IRE1αin epithelial tissues of OVAchallenged mice
(Park et al., 2015).
Chung and colleagues have recently explored the antiasthmatic
role of kaempferol through multiple mechanisms such as lowering
the elevated inflammatory cell numbers, inhibiting the enhancement
of TNFαprotein levels and Th2 cytokines (IL4, IL5, and IL13), and
suppressing the phosphorylation of Akt and eosinophilia. Moreover,
kaempferol blocks the total immunoglobulin (Ig) E levels in the serum
and bronchoalveolar lavage fluid (Chung et al., 2015). Furthermore,
kaempferol affects protection from the human Paraquat (PQ)exposed
bronchial epithelium BEAS2B cells by inhibiting mucin gene expres-
sion via NFκB. It additionally suppresses the PQ phosphorylation of
ERK and cJNK (Podder, Song, Song, & Kim, 2014). In BEAS2B cells
of BALB/c mice, kaempferol suppressed the lipopolysaccharide
(LPS)induced bronchial EMT. It exhibited a significant inhibitory effect
on the TGFβinduced EMT process by reversing Ecadherin expres-
sion and by retarding the induction of Ncadherin and αsmooth mus-
cle actin (αSMA). Kaempferol also inhibited epithelial excrescency,
collagen deposition, and goblet hyperplasia, and it suppressed TGFβ
entailed epithelial proteaseactivated receptor1 (Gong et al., 2014).
A study conducted by Gong et al. (2013) revealed that administra-
tion of kaempferol at less than or equal to 20 μM in BALB/c mice sen-
sitized with OVA inhibited LPSinduced IL8 production through the
TLR4 activation, suppressed eotaxin1 induction, attenuated the
upregulated CXCR2 expression and macrophage inflammatory
protein2 production, and allayed the airway tissue levels of eotaxin
1 and eotaxin receptor CCR3 in a dosedependent manner. It addition-
ally suppressed the IL8inflamed Tyk2 activation and prevented acti-
vation of the STAT1/3 signaling concomitant and downregulated the
expression of Tykinhibiting SOCS3 (Gong et al., 2013). On the
other hand, Medeiros et al. (2009) reported that kaempferol, in
BALB/c mice ovalbumin (OVA)/alum, dosedependently lowered the
total leukocyte and eosinophil counts in bronchoalveolar lavage fluid.
It also lowered the B220(+), CD4(+), MHC class II, CD40 molecule
expressions, inhibited mucus production, and ameliorated the airway
hyper responsiveness (AHR). Moreover, it impaired Th2 cytokine pro-
duction (IL5 and IL13) and did not induce a Th1 pattern of inflamma-
tion (Medeiros et al., 2009).
2.8 |Antiplatelet aggregation capacity
In animal models, including thrombininduced acute thromboembo-
lism, collagen/epinephrine, and FeCl
3
induced carotid arterial throm-
bus models, kaempferol markedly suppressed enzymatic activities of
thrombin, FXaas, and fibrin polymer formation. Similarly, kaempferol
attenuated the phosphorylation of ERK1/2, p38, cJNK1/2, and
phosphoinositide 3kinase (PI3K)/PKB (AKT) and protected against
FIGURE 3 Antiinflammatory mechanism associated with kaempferol [Colour figure can be viewed at wileyonlinelibrary.com]
8IMRAN ET AL.
thrombosis development (Choi et al., 2015). In vitro and in prolonged
in vivo thrombotic response in carotid arteries of mice, kaempferol
caused inhibited production of superoxide anion stimulated by colla-
gen, suppressed NOX activation and collageninduced phosphoryla-
tion of p47 (phox), and led to ROSdependent inactivation of SH2
domaincontaining protein tyrosine phosphatase2 (SHP2). It also
suppressed the specific tyrosine phosphorylation of key components
(Vav1, Syk, PLCγ2, and Btk) of collagen receptor signaling pathways,
and it attenuated the downstream responses, including Pselectin
surface exposure, cytosolic calcium elevation, and integrinαIIbβ3
activation (Wang et al., 2015).
2.9 |Bone health promoting capacity
Effect of kaempferol on bone marrowderived mesenchymal stem
cells differentiation and inflammation has been recently investigated.
Results revealed that kaempferol improves cell viability and inhibits
cell apoptosis and cell proliferation induced by lipopolysaccharide
(LPS). In addition, kaempferol elevated the LPSinduced level of
chondrogenic markers (SOX9, collagen II and aggrecan), and lowered
the level of matrixdegrading enzymes,i.e., matrix metalloprotease
(MMP)3 and MMP13. These results highlight the protective effects
of kaempferol against osteoporosis and obesity (Zhu et al., 2017).
Kim et al. (2016) showed that kaempferol (10 μM) promotes the pro-
liferation, differentiation, and mineralization of osteoblasts. It addi-
tionally increased the expression of the osteoblastactivated factors
RUNX2, BMP2, osterix, collagen I, induced autophagy, and enhanced
levels of the autophagyrelated factors beclin1, the conversion of
LC3II from LC3I, and SQSTM1/p62 (Kim et al., 2016). It also sup-
presses the proliferation and migration of VSMC by modulating the
microRNA expression levels and BMP4 signaling pathway, activating
the BMP signaling pathway, inducing miR21 expression, and down
regulating DOCK4, 5, and 7 (Kim et al., 2015). On the other hand,
research findings showed that kaempferol inhibits IL1βstimulated
RANKLmediated osteoclast differentiation. Additionally it suppresses
IL1βstimulated RANKLmediated phosphorylation of ERK 1/2, p38,
and JNK MAP kinases, and suppresses the expressions of NFATc1
and cFos (Lee, Lee, Sung, & Yoo, 2014).
Release of inflammatory cytokines such as MMP1, MMP3, and
COX2 from rheumatoid arthritis synovial fibroblasts is associated
with articular bone and cartilage destruction. On the other hand,
kaempferol significantly inhibits production of nitric oxide synthase
and COX2 enzymes, proliferation of both unstimulated and IL1β
stimulated RASFs, formation of prostaglandin E2 (PGE2) as well as
mRNA and protein expression of MMP1, and MMP3. Moreover,
kaempferol reduced the phosphorylation of p38, ERK1/2,& JNK,
and activation of NFκB induced by IL1β(Gupta et al., 2013). Several
studies conducted by different investigators reported that utilization
of kaempferol (5.0 μM) against primary rat calvarial osteoblasts
enhanced the alkaline phosphatase activity and mineralization of the
cells. These studies included cytoskeletal proteins, chaperone extracel-
lular matrix protein, intracellular signaling protein, and proteins
involved in glycolysis and cellmatrix interactions. In addition, it up
regulated the HSP70 and cytokeratin14 levels, and downregulated
the aldose reductase and caldesmon expression. In osteoblasts,
kaempferol altered cellular metabolism by slowing the polyol pathway
(Trivedi et al., 2009).
Trivedi and colleagues investigated the antiosteoclastogenic role
of kaempferol. They found that kaempferol in rat primary osteoblasts
markedly attenuates adipocyte formation and shows lower serum ALP
(bone turnover marker) and higher bone mineral density (BMD) in the
trabecular regions (proximal tibia, femur neck, and vertebrae) (Trivedi
et al., 2008). Furthermore, in MG63 cultured human osteoblasts,
kaempferol significantly enhanced the activity of alkaline phosphatase.
This process involves activation of the extracellular regulated kinase
(ERK) pathway, and activation of the estrogen receptor (ER) (Prouillet
et al., 2004).
FIGURE 4 Metabolic effect of kaempferol
indicating the mechanism associated with its
antidiabetic, antiobesity, cardiovascular and
oxidative damage protective perspective.
[Colour figure can be viewed at
wileyonlinelibrary.com]
IMRAN ET AL.9
2.10 |Antiobesity effect
Research by Lee and colleagues showed that kaempferol inhibits lipid
accumulation in adipocytes and zebrafish. It decreases the expression
levels of LPAATθ(lysophosphatidic acid acyltransferase), SREBP1C
(fatty acid synthetic proteins) lipin1, DGAT1 (triglyceride synthetic
enzymes), and FASN. Results also revealed that kaempferol delays cell
cycle progression from the S to G2/M phase, blocks mammalian target
of rapamycin (mTOR) and phosphorylation of AKT (protein kinase B),
and regulates cyclins in a dosedependent manner. In addition,
kaempferol downregulates the CCAATenhancer binding proteins β
(C/EBPβ) and proearly adipogenic factors such as Krüppellike factors
(KLFs) 4 and 5, and upregulates the antiearly adipogenic factors
(pref1(preadipocyte factor1), and KLF2). Moreover, attenuation of
late adipogenic factors such as C/EBPαand peroxisome proliferator
activated receptor γ(PPARγ) by kaempferol was reported (Lee et al.,
2015). On the other hand, a group of researchers confirmed the find-
ings that kaempferol plays an inhibitory role against 3T3L1 adipo-
cytes through multiple mechanisms such as (a) reduction of the
levels of adipogenic transcription factors (Cebpβ, Pparγ, Rxrβ, Srebp1,
Rorα, Lxrβ) and genes involved in triglyceride biosynthesis (Agpat2,
Gpd1, Dgat2) and (b) enhancement of lipolysisrelated genes (Lsr,
Tnfα, and Cel). Furthermore, kaempferol considerably repressed the
rosiglitazoneinduced PPARγtranscriptional activity (Park et al.,
2012). Displayed in Figure 4 is the metabolic effect of kaempferol indi-
cating the mechanism associated with its antidiabetic, antiobesity,
cardiovascular, and oxidative damage protective perspectives.
3|CONCLUSIONS
Chemopreventive and chemotherapeutic roles of compounds derived
from natural sources (plants and animals) have been gaining popularity
and attracting the attention of medicinal chemists, pharmacologists,
and dieticians due to the many benefits of these substances to human
health. These naturally derived compounds can also be used in the
fight against diseases such as cardiovascular disorders, cancer insur-
gence, and immune dysfunction, diabetes, oxidative stress, among
others. In addition, conventional therapies such as the use of natural
products in treating diseases, has attracted the attention of the scien-
tific and medical communities due to their lesser side effects and cost
in comparison with synthetic drugs. Kaempferol, a widely used dietary
flavonoid found in plant foods, such as tea, cabbage, broccoli, endive,
kale, beans, tomato, strawberries, leek, and grapes is considered as
one of the most abundant antioxidants in the human diet, and plays
a significant role in scavenging radicals and inflammation. Its efficacy,
a broad range of activity, and low toxicity compared with other exam-
ined compounds, make it an attractive chemical in the fight against
diseases (including cancer). In this review, we have shown through
documented research that kaempferol offers a widespread range of
preventive and therapeutic options against several diseases, along
with a description of the various mechanisms by which this compound
exerts its action. In short, this review reveals that kaempferol can be
an important complementary medicine for the prevention and treat-
ment of different illnesses, owing to its natural origin, safety, and
low cost relative to synthetic drugs. However, further studies (includ-
ing clinical trials on humans) of this natural compound are needed to
establish its efficacy and safety.
CONFLICT OF INTEREST
Authors declare no conflict of interest.
ORCID
Mohammad S. Mubarak http://orcid.org/0000-0002-9782-0835
REFERENCES
Alkhalidy, H., Moore, W., Zhang, Y., McMillan, R., Wang, A., Ali, M.,
Liu, D. (2015). Small molecule kaempferol promotes insulin sensitivity
and preserved pancreatic βcell mass in middleaged obese diabetic
mice. Journal of Diabetes Research,2015.
AlNumair, K. S., Chandramohan, G., Veeramani, C., & Alsaif, M. A. (2015).
Ameliorative effect of kaempferol, a flavonoid, on oxidative stress in
streptozotocininduced diabetic rats. Redox Report,20(5), 198209.
AlNumair, K. S., Veeramani, C., Alsaif, M. A., & Chandramohan, G. (2015).
Influence of kaempferol, a flavonoid compound, on membranebound
ATPases in streptozotocininduced diabetic rats. Pharmaceutical Biol-
ogy,53(9), 13721378.
Arts, I. C., & Hollman, P. C. (2005). Polyphenols and disease risk in epidemi-
ologic studies. The American Journal of Clinical Nutrition,81,
317S325S.
Azevedo, C., CorreiaBranco, A., Araújo, J. R., Guimarães, J. T., Keating, E.,
& Martel, F. (2015). The chemopreventive effect of the dietary com-
pound kaempferol on the MCF7 human breast cancer cell line is
dependent on inhibition of glucose cellular uptake. Nutrition and Can-
cer,67(3), 504513.
CalderónMontaño, J. M., BurgosMorón, E., PérezGuerrero, C., & López
Lázaro, M. (2011). A review on the dietary flavonoid kaempferol. Mini
Reviews in Medicinal Chemistry,11(4), 298340.
Chang, S. C., Cassidy, A., Willett, W. C., Rimm, E. B., O'Reilly, E. J., &
Okereke, O. I. (2016). Dietary flavonoid intake and risk of incident
depression in midlife and older women. The American Journal of Clinical
Nutrition,104(3), 704714.
Cho, H. J., & Park, J. H. Y. (2013). Kaempferol induces cell cycle arrest in
HT29 human colon cancer cells. Journal of Cancer Prevention,18(3),
257263.
Choi, J. H., Park, S. E., Kim, S. J., & Kim, S. (2015). Kaempferol inhibits
thrombosis and platelet activation. Biochimie,115, 177186.
Choi, Y. J., Lee, Y. H., & Lee, S. T. (2015). Galangin and kaempferol suppress
phorbol12myristate13acetateinduced matrix metalloproteinase9
expression in human fibrosarcoma HT1080 cells. Molecular Cell,
38(2), 151155.
Chung, M. J., Pandey, R. P., Choi, J. W., Sohng, J. K., Choi, D. J., & Park, Y.
(2015). Inhibitory effects of kaempferol3Orhamnoside on
ovalbumininduced lung inflammation in a mouse model of allergic
asthma. International Immunopharmacology,25(2), 302310.
Dang, Q., Song, W., Xu, D., Ma, Y., Li, F., Zeng, J., Li, L. (2015).
Kaempferol suppresses bladder cancer tumor growth by inhibiting cell
proliferation and inducing apoptosis. Molecular Carcinogenesis,54(9),
831840.
Das, A., Majumder, D., & Saha, C. (2017). Correlation of binding efficacies
of DNA to flavonoids and their induced cellular damage. Journal of
Photochemistry and Photobiology B: Biology,170, 256262.
Feng, H., Cao, J., Zhang, G., & Wang, Y. (2017). Kaempferol attenuates car-
diac hypertrophy via regulation of ASK1/MAPK signaling pathway and
oxidative stress. Planta Medica,83(10), 837845.
Gao, S. S., Choi, B.M., Chen, X. Y., Zhu, R. Z., Kim, Y., So, H., Kim, B. R.
(2010). Kaempferol suppresses cisplatininduced apoptosis via
10 IMRAN ET AL.
inductions of heme oxygenase1 and glutamatecysteine ligase cata-
lytic subunit in HEIOC1 cells. Pharmaceutical Research,27(2),
235245.
Ghosh, D., Dey, S. K., & Saha, C. (2014). Antagonistic effects of black tea
against gamma radiationinduced oxidative damage to normal lympho-
cytes in comparison with cancerous K562 cells. Radiation and
Environmental Biophysics,53(4), 695704.
Godos, J., Castellano, S., Ray, S., Grosso, G., & Galvano, F. (2018). Dietary
polyphenol intake and depression: Results from the Mediterranean
healthy eating, lifestyle and aging (MEAL) study. Molecules,23(5).
Gong, J. H., Cho, I. H., Shin, D., Han, S. Y., Park, S.H., & Kang, Y. H. (2014).
Inhibition of airway epithelialtomesenchymal transition and fibrosis
by kaempferol in endotoxininduced epithelial cells and ovalbumin
sensitized mice. Laboratory Investigation,94(3), 297308.
Gong, J. H., Shin, D., Han, S. Y., Kim, J. L., & Kang, Y. H. (2012). Kaempferol
suppresses eosionphil infiltration and airway inflammation in airway
epithelial cells and in mice with allergic asthma. The Journal of Nutrition,
142(1), 4756.
Gong, J. H., Shin, D., Han, S. Y., Park, S. H., Kang, M. K., Kim, J. L., & Kang,
Y. H. (2013). Blockade of airway inflammation by kaempferol via
disturbing TykSTAT signaling in airway epithelial cells and in asthmatic
mice. Evidencebased Complementary and Alternative Medicine, 2013.
Grosso, G., Godos, J., LamuelaRaventos, R., Ray, S., Micek, A., Pajak, A.,
Galvano, F. (2017). A comprehensive metaanalysis on dietary flavo-
noid and lignan intake and cancer risk: Level of evidence and
limitations. Molecular Nutrition & Food Research,61(4), 1600930.
Grosso, G., Micek, A., Godos, J., Pajak, A., Sciacca, S., Galvano, F., &
Giovannucci, E. L. (2017). Dietary flavonoid and lignan intake and mor-
tality in prospective cohort studies: Systematic review and dose
response metaanalysis. American Journal of Epidemiology,185(12),
13041316.
Guo, Z., Liao, Z., Huang, L., Liu, D., Yin, D., & He, M. (2015). Kaempferol
protects cardiomyocytes against anoxia/reoxygenation injury via mito-
chondrial pathway mediated by SIRT1. European Journal of
Pharmacology,761, 245253.
Gupta, G. K., Kumar, A., Khedgikar, V., Gautam, J., Nagar, G. K., Gupta, V.,
Mishra, P. R. (2013). Osteogenic efficacy enhancement of kaempferol
through an engineered layerbylayer matrix: A study in ovariectomized
rats. Nanomedicine,8(5), 757771.
Haidari, F., Keshavarz, S. A., Shahi, M. M., Mahboob, S. A., & Rashidi, M. R.
(2011). Effects of parsley (Petroselinum crispum) and its flavonol constit-
uents, kaempferol and quercetin, on serum uric acid levels, biomarkers
of oxidative stress and liver Xanthine oxidoreductase aactivity
inOxonateinduced hyperuricemic rats. Iranian Journal of Pharmaceuti-
cal Research,10(4), 811819.
Halimah, E., Diantini, A., Destiani, D. P., Destiani, I. S., Pradipta, H. S.,
Sastramihardja, K. L., Hiroshi, K. (2015). Induction of caspase cascade
pathway by kaempferol3Orhamnoside in LNCaP prostate cancer cell
lines. Biomedical Reports,3(1), 115117.
Heijnen, C., Haenen, G., Van Acker, F., Van der Vijgh, W., & Bast, A. (2001).
Flavonoids as peroxynitrite scavengers: The role of the hydroxyl
groups. Toxicology In Vitro,15(1), 36.
Heim, K. E., Tagliaferro, A. R., & Bobilya, D. J. (2002). Flavonoid antioxi-
dants: Chemistry, metabolism and structureactivity relationships.
Journal of Nutritional Biochemistry,13(10), 572584.
Hooper, L., Kroon, P. A., Rimm, E. B., Cohn, J. S., Harvey, I., Le Cornu, K. A.,
Cassidy, A. (2008). Flavonoids, flavonoidrich foods, and cardiovas-
cular risk: A metaanalysis of randomized controlled trials.
The American Journal of Clinical Nutrition,88(1), 3850.
Huang, W. W., Tsai, S. C., Peng, S. F., Lin, M. W., Chiang, J. H., Chiu, Y. J.,
Yang, J. S. (2013). Kaempferol induces autophagy through AMPK and
AKT signaling molecules and causes G2/M arrest via downregulation
of CDK1/cyclin B in SKHEP1 human hepatic cancer cells. Interna-
tional Journal of Oncology,42(6), 20692077.
Jeong, H. Y., Sung, G. H., Kim, J. H., Yoon, J. Y., Yang, Y., Park, J. G., Cho,
J. Y. (2014). Syk and Src are major pharmacological targets of a Cerbera
manghas methanol extract with kaempferolbased antiinflammatory
activity. Journal of Ethnopharmacology,151(2), 960969.
Jo, E., Park, S. J., Choi, Y. S., Jeon, W. K., & Kim, B. C. (2015). Kaempferol
suppresses transforming growth factorβ1induced epithelialto
mesenchymal transition and migration of A549 lung cancer cells by
inhibiting Akt1mediated phosphorylation of Smad3 at threonine179.
Neoplasia,17(7), 525537.
Jose, J., Sudhakaran, S., Kumar, S., Jayaraman, S., & Variyar, E. J. (2014). A
comparative evaluation of anticancer activities of flavonoids isolated
from Mimosa pudica,Aloe vera and Phyllanthus niruri against human
breast carcinoma cell line (MCF7) using MTT assay. International
Journal of Pharmacy and Pharmaceutical Sciences,6(2), 319322.
Kang, J. W., Kim, J. H., Song, K., Kim, S. H., Yoon, J. H., & Kim, K. S. (2010).
Kaempferol and quercetin, components of Ginkgo biloba extract (EGb
761), induce caspase3dependent apoptosis in oral cavity cancer cells.
Phytotherapy Research,24(1), 213217.
Kashafi, E., Moradzadeh, M., Mohamadkhani, A., & Erfanian, S. (2017).
Kaempferol increases apoptosis in human cervical cancer HeLa cells
via PI3K/AKT and telomerase pathways. Biomedicine and Pharmaco-
therapy,89, 573577.
Kim, I. R., Kim, S. E., Baek, H. S., Kim, B. J., Kim, C. H., Chung, I. K., Shin, S.
H. (2016). The role of kaempferolinduced autophagy on differentia-
tion and mineralization of osteoblastic MC3T3E1 cells. Evidence
based Complementary and Alternative Medicine,16(1), 333.
Kim, K., Kim, S., Moh, S. H., & Kang, H. (2015). Kaempferol inhibits vascular
smooth muscle cell migration by modulating BMPmediated miR21
expression. Molecular and Cellular Biochemistry,407(12), 143149.
Kim, S. H., Hwang, K. A., & Choi, K. C. (2016). Treatment with kaempferol
suppresses breast cancer cell growth caused by estrogen and triclosan
in cellular and xenograft breast cancer models. The Journal of
Nutritional Biochemistry,28,7082.
Kim, S. H., Park, J. G., Lee, J., Yang, W. S., Park, G. W., Kim, H. G., Cho, J.
Y. (2015). The dietary flavonoid Kaempferol mediates anti
inflammatory responses via the Src, Syk, IRAK1, and IRAK4 molecular
targets. Mediators of Inflammation,2015, Article ID 904142.
Klaunig, J. E., & Kamendulis, L. M. (2004). The role of oxidative stress in
carcinogenesis. Annual Review of Pharmacology and Toxicology,44,
239267.
Kumar, S., & Pandey, A. K. (2013). Phenolic content, reducing power and
membrane protective activities of Solanum xanthocarpum root
extracts. VEGETOS: An International Journal of Plant Research,26(1),
301307.
Lagoa, R., LopezSanchez, C., SamhanArias, A. K., Gañan, C. M., Garcia
Martinez, V., & GutierrezMerino, C. (2009). Kaempferol protects
against rat striatal degeneration induced by 3nitropropionic acid.
Journal of Neurochemistry,111(2), 473487.
Lee, C. F., Yang, J. S., Tsai, F. J., Chiang, N. N., Lu, C. C., Huang, Y. S.,
Chen, F. A. (2016). Kaempferol induces ATM/p53mediated death
receptor and mitochondrial apoptosis in human umbilical vein endothe-
lial cells. International Journal of Oncology,48(5), 20072014.
Lee, G. A., Choi, K. C., & Hwang, K. A. (2017). Kaempferol, a phytoestrogen,
suppressed triclosaninduced epithelialmesenchymal transition and
metastaticrelated behaviors of MCF7 breast cancer cells.
Environomental Toxicology and Pharmacology,49,48
57.
Lee, H. S., Cho, H. J., Yu, R., Lee, K. W., Chun, H. S., & Park, J. H. Y. (2014).
Mechanisms underlying apoptosisinducing effects of Kaempferol in
HT29 human colon cancer cells. International Journal of Molecular
Sciences,15(2), 27222737.
Lee, J., & Kim, J. H. (2016). Kaempferol inhibits pancreatic cancer cell
growth and migration through the blockade of EGFRrelated pathway
in vitro. PLoS One,11(5), e0155264.
Lee, W. S., Lee, E. G., Sung, M. S., & Yoo, W. H. (2014). Kaempferol inhibits
IL1βstimulated, RANKLmediated osteoclastogenesis via downregu-
lation of MAPKs, cFos, and NFATc1. Inflammation,37, 12211230.
IMRAN ET AL.11
Lee, Y. J., Choi, H. S., Seo, M. J., Jeon, H. J., Kim, K. J., & Lee, B. Y. (2015).
Kaempferol suppresses lipid accumulation by inhibiting early adipogen-
esis in 3T3L1 cells and zebrafish. Food & Function,6(8), 28242833.
Lee, Y. J., Suh, K. S., Choi, M. C., Chon, S., Oh, S., Woo, J. T., Kim, Y. S.
(2010). Kaempferol protects HITT15 pancreatic beta cells from 2
deoxyDriboseinduced oxidative damage. Phytotherapy Research,
24(3), 419423.
Li, S., & Pu, X. P. (2011). Neuroprotective effect of kaempferol against a
1methyl4phenyl1, 2, 3, 6tetrahydropyridineinduced mouse model
of Parkinson's disease. Biological and Pharmaceutical Bulletin,34(8),
12911296.
Liao, W., Chen, L., Ma, X., Jiao, R., Li, X., & Wang, Y. (2016). Protective
effects of kaempferol against reactive oxygen speciesinduced hemoly-
sis and its antiproliferative activity on human cancer cells. European
Journal of Medicinal Chemistry,114,2432.
Lin, F., Luo, X., Tsun, A., Li, Z., Li, D., & Li, B. (2015). Kaempferol enhances
the suppressive function of Treg cells by inhibiting FOXP3 phosphory-
lation. International Immunopharmacology,28(2), 859865.
Liu, Y. J., Zhan, J., Liu, X. L., Wang, Y., Ji, J., & He, Q. Q. (2014). Dietary fla-
vonoids intake and risk of type 2 diabetes: A metaanalysis of
prospective cohort studies. Clinical Nutrition,33(1), 5963.
LópezSánchez, C., MartínRomero, F. J., Sun, F., Luis, L., SamhanArias, A.
K., GarcíaMartíne, V., & GutiérrezMerino, C. (2007). Blood micromo-
lar concentrations of kaempferol afford protection against
ischemia/reperfusioninduced damage in rat brain. Brain Research,
1182, 123137.
Luo, C., Yang, H., Tang, C., Yao, G., Kong, L., He, H., & Zhou, Y. (2015).
Kaempferol alleviates insulin resistance via hepatic IKK/NFκB signal
in type 2 diabetic rats. International Immunopharmacology,28(1),
744750.
Mahomoodally, M. F., GuribFakim, A., & Subratty, A. H. (2005). Antimicro-
bial activities and phytochemical profiles of endemic medicinal plants
of Mauritius. Pharmaceutical Biology,43(3), 237242.
Medeiros, K., Faustino, L., Borduchi, E., Nascimento, R. J., Silva, T. M.,
Gomes, E., Russo, M. (2009). Preventive and curative glycoside
kaempferol treatments attenuate the TH2driven allergic airway dis-
ease. International Immunopharmacology,9(13), 15401548.
Moon, Y. J., Wang, X., & Morris, M. E. (2006). Dietary flavonoids: Effects
on xenobiotic and carcinogen metabolism. Toxicology In Vitro,20,
187210.
Nam, S.Y., Jeong, H.J., & Kim, H.M. (2017). Kaempferol impedes IL32
induced monocytemacrophage differentiation. ChemicoBiological
Interactions,274, 107115.
Pal, S., Dey, S. K., & Saha, C. (2014). Inhibition of catalase by tea catechins
in free and cellular state: A biophysical approach. PLoS One,9(7),
e102460.
Park, S. H., Gong, J. H., Choi, Y. J., Kang, M. K., Kim, Y. H., & Kang, Y. H.
(2015). Kaempferol inhibits endoplasmic reticulum stressassociated
mucus hypersecretion in airway epithelial cells and ovalbumin
sensitized mice. PLoS One,10(11), 143526.
Park, U. H., Jeong, J. C., Jang, J. S., Sung, M. R., Youn, H., Lee, S. J., Um, S.
J. (2012). Negative regulation of adipogenesis by kaempferol, a compo-
nent of Rhizoma Polygonati falcatum in 3T3L1 cells. Biological and
Pharmaceutical Bulletin,35(9), 15251533.
Peng, X., Zhang, G., Liao, Y., & Gong, D. (2016). Inhibitory kinetics and
mechanism of kaempferol on αglucosidase. Food Chemistry,190,
207215.
Podder, B., Song, K. S., Song, H. Y., & Kim, Y. S. (2014). Cytoprotective
effect of kaempferol on paraquatexposed BEAS2B cells via modulat-
ing expression of MUC5AC. Biological and Pharmaceutical Bulletin,
37(9), 14861494.
Prouillet, C., Mazière, J. C., Maziere, C., Wattel, A., Brazier, M., & Kamel, S.
(2004). Stimulatory effect of naturally occurring flavonols quercetin
and kaempferol on alkaline phosphatase activity in MG63 human
osteoblasts through ERK and estrogen receptor pathway. Biochemical
Pharmacology,67(7), 13071313.
Qin, Y., Cui, W., Yang, X., & Tong, B. (2016). Kaempferol inhibits the
growth and metastasis of cholangiocarcinoma in vitro and in vivo. Acta
Biochimica et Biophysica Sinica,48(3), 238245.
Qiu, W., Lin, J., Zhu, Y., Zeng, L., Su, M., & T ian, Y. (2017). Kaempferol mod-
ulates DNA methylation and downregulates DNMT3B in bladder
cancer. Cellular Physiology and Biochemistry,41(4), 13251335.
Qu, W., Fan, L., Kim, Y. C., Ishikawa, S., IguchiAriga, S. M., Pu, X. P., &
Ariga, H. (2009). Kaempferol derivatives prevent oxidative stress
induced cell death in a DJ1dependent manner. Journal of Pharmaco-
logical Sciences,110(2), 191200.
Roth, A., Schaffner, W., & Hertel, C. (1999). Phytoestrogen kaempferol
(3,4,5,7tetrahydroxyflavone) protects PC12 and T47D cells from
βamyloidinduced toxicity. Journal of Neuroscience Research,57(3),
399404.
Saw, C. L. L., Guo, Y., Yang, A. Y., ParedesGonzalez, X., Ramirez, C., Pung,
D., & Kong, A. N. (2014). The berry constituents quercetin, kaempferol,
and pterostilbene synergistically attenuate reactive oxygen species:
Involvement of the Nrf2ARE signaling pathway. Food and Chemical
Toxicology,72, 303311.
Scalbert, A., & Williamson, G. (2000). Dietary intake and bioavailability of
polyphenols. The Journal of Nutrition,130, 2073S2085S.
Shakya, G., Manjini, S., Hoda, M., & Rajagopalan, R. (2014). Hepatoprotec-
tive role of kaempferol during alcoholand ΔPUFAinduced oxidative
stress. Journal of Basic and Clinical Physiology and Pharmacology,25(1),
7379.
Sharoar, M. G., Thapa, A., Shahnawaz, M., Ramasamy, V. S., Woo, E. R.,
Shin, S. Y., & Park, I. S. (2012). Keampferol3Orhamnoside abrogates
amyloid beta toxicity by modulating monomers and remodeling oligo-
mers and fibrils to nontoxic aggregates. Journal of Biomedical Science,
19(1), 104.
Shin, D., Park, S. H., Choi, Y. J., Kim, Y. H., Antika, L. D., Habibah, N. U.,
Kang, Y. H. (2015). Dietary compound Kaempferol inhibits airway
thickening induced by allergic reaction in a bovine serum albumin
induced model of asthma. International Journal of Molecular Sciences,
16(12), 2998029995.
Song, H., Bao, J., Wei, Y., Chen, Y., Mao, X., Li, J., Xue, Y. (2015).
Kaempferol inhibits gastric cancer tumor growth: An in vitro and
in vivo study. Oncology Reports,33(2), 868874.
Song, W., Dang, Q., Xu, D., Chen, Y., Zhu, G., Wu, K., Li, L. (2014).
Kaempferol induces cell cycle arrest and apoptosis in renal cell carci-
noma through EGFR/p38 signaling. Oncology Reports,31(3),
13501356.
Suchal, K., Malik, S., Gamad, N., Malhotra, R. K., Goyal, S. N., Chaudhary, U.,
Arya, D. S. (2016). Kaempferol attenuates myocardial ischemic injury
via inhibition of MAPK signaling pathway in experimental model of
myocardial ischemiareperfusion injury. Oxidative Medicine and Cellular
Longevity,2016. 7580731
Suchal, K., Malik, S., Gamad, N., Malhotra, R. K., Goyal, S. N., Bhatia, J., &
Arya, D. S. (2016a). Kampeferol protects against oxidative stress and
apoptotic damage in experimental model of isoproterenolinduced car-
diac toxicity in rats. Phytomedicine,23(12), 14011408.
Suchal, K., Malik, S., Khan, S. I., Malhotra, R. K., Goyal, S. N., Bhatia, J.,
Arya, D.S. (2017). Molecular pathways involved in the amelioration
of myocardial injury in diabetic rats by kaempferol. International Journal
of Molecular Sciences,18(5). https://doi.org/10.3390/ijms18051001
Swanson, H. I., Choi, E. Y., Helton, W. B., Gairola, C. G., & Valentino, J.
(2014). Impact of apigenin and kaempferol on human head and neck
squamous cell carcinoma. Oral Surgery, Oral Medicine, Oral Pathology,
Oral Radiology,117(2), 214220.
Tang, X. l., Liu, J. X., Dong, W., Li, P., Li, L., Hou, J. C., Ren, J. G. (2015).
Protective effect of kaempferol on LPS plus ATPinduced inflammatory
response in cardiac fibroblasts. Inflammation,38(1), 94101.
Trivedi, R., Kumar, A., Gupta, V., Kumar, S., Nagar, G. K., Romero, J. R.,
Chattopadhyay, N. (2009). Effects of Egb 761 on bone mineral density,
bone microstructure, and osteoblast function: Possible roles of querce-
tin and kaempferol. Molecular and Cellular Endocrinology,302(1), 8691.
12 IMRAN ET AL.
Trivedi, R., Kumar, S., Kumar, A., Siddiqui, J. A., Swarnkar, G., Gupta, V.,
Chattopadhyay, N. (2008). Kaempferol has osteogenic effect in ovari-
ectomized adult SpragueDawley rats. Molecular and Cellular
Endocrinology,289(12), 8593.
Tu, L. Y., Bai, H. H., Cai, J. Y., & Deng, S. P. (2016). The mechanism of
kaempferol induced apoptosis and inhibited proliferation in human cer-
vical cancer SiHa cell: From macro to nano. Scanning,38(6), 644653.
Varshney, R., Gupta, S., & Roy, P. (2017). Cytoprotective effect of
kaempferol against palmitic acidinduced pancreatic βcell death
through modulation of autophagy via AMPK/mTOR signaling pathway.
Molecular and Cellular Endocrinology,448,120.
Wang, L., Lee, I. M., Zhang, S. M., Blumberg, J. B., Buring, J. E., & Sesso, H.
D. (2009). Dietary intake of selected flavonols, flavones, and flavonoid
rich foods and risk of cancer in middleaged and older women.
The American Journal of Clinical Nutrition,89(3), 905912.
Wang, L., Tu, Y. C., Lian, T.W., Hung, J. T., Yen, J. H., & Wu, M. J. (2006).
Distinctive antioxidant and antiinflammatory effects of flavonols.
Journal of Agricultural and Food Chemistry,54(26), 97989804.
Wang, M., Sun, J., Jiang, Z., Xie, W., & Zhang, X. (2015). Hepatoprotective
effect of kaempferol against alcoholic liver injury in mice. The American
Journal of Chinese Medicine,43, 241254.
Wang, S. B., Jang, J. Y., Chae, Y. H., Min, J. H., Baek, J. Y., Kim, M., Chang,
T. S. (2015). Kaempferol suppresses collageninduced platelet activa-
tion by inhibiting NADPH oxidase and protecting SHP2 from
oxidative inactivation. Free Radical Biology & Medicine,83,4153.
Wu, L. Y., Lu, H. F., Chou, Y. C., Shih, Y. L., Bau, D. T., Chen, J. C., Chung,
J. G. (2015). Kaempferol induces DNA damage and inhibits DNA repair
associated protein expressions in human promyelocytic leukemia
HL60 cells. The American Journal of Chinese Medicine,43, 365382.
Xiao, H. B., Lu, X. Y., Liu, Z. K., & Luo, Z. F. (2016). Kaempferol inhibits the
production of ROS to modulate OPNαvβ3 integrin pathway in
HUVECs. Journal of Physiology and Biochemistry,72(2), 303313.
Xu, X., Zhao, C., Peng, Q., Xie, P., & Liu, Q. (2017). Kaempferol inhibited
VEGF and PGF expression and in vitro angiogenesis of HRECs under
diabeticlike environment. Brazilian Journal of Medical and Biological
Research,50(3), e5396.
Xu, Y. C., Leung, S. W., Leung, G. P., & Man, R. Y. (2015). Kaempferol
enhances endotheliumdependent relaxation in the porcine coronary
artery through activation of largeconductance Ca(2+) activated K(+)
channels. British Journal of Pharmacology,172(12), 30033014.
Yang, E. J., Kim, G. S., Jun, M., & Song, K. S. (2014). Kaempferol attenuates
the glutamateinduced oxidative stress in mousederived hippocampal
neuronal HT22 cells. Food & Function,5(7), 13951402.
Yao, K., Chen, H., Liu, K., Langfald, A., Yang, G., Zhang, Y., Dong, Z.
(2014). Kaempferol targets RSK2 and MSK1 to suppress UV
radiationinduced skin cancer. Cancer Prevention Research,7(9),
958967.
Zang, Y., Zhang, L., Igarashi, K., & Yu, C. (2015). The antiobesity and anti
diabetic effects of kaempferol glycosides from unripe soybean leaves in
highfatdiet mice. Food & Function,6(3), 834841.
Zhang, Y., Zhen, W., Maechler, P., & Liu, D. (2013). Small molecule
kaempferol modulates PDX1 protein expression and subsequently
promotes pancreatic βcell survival and function via CREB. J Nutr
Biochem,24(4), 638646.
Zhang, R., Ai, X., Duan, Y., Xue, M., He, W., Wang, C., Liu, H. (2017).
Kaempferol ameliorates H9N2 swine influenza virusinduced acute
lung injury by inactivation of TLR4/MyD88mediated NFκB and
MAPK signaling pathways. Biomedicine & Pharmacotherapy,89,
660672.
Zhou, M., Ren, H., Han, J., Wang, W., Zheng, Q., & Wang, D. (2015).
Protective effects of kaempferol against myocardial
ischemia/reperfusion injury in isolated rat heart via antioxidant activity
and inhibition of glycogen synthase kinase3. Oxidative Medicine and
Cellular Longevity,2015.
Zhu, J., Tang, H., Zhang, Z., Zhang, Y., Qiu, C., Zhang, L., Li, F. (2017).
Kaempferol slows intervertebral disc degeneration by modifying
LPSinduced osteogenesis/adipogenesis imbalance and inflammation
response in BMSCs. International Immunopharmacology,43, 236242.
How to cite this article: Imran M, Rauf A, Shah ZA, et al.
Chemopreventive and therapeutic effect of the dietary flavo-
noid kaempferol: A comprehensive review. Phytotherapy
Research. 2018;113. https://doi.org/10.1002/ptr.6227
IMRAN ET AL.13
... Its derivatives include kaempferol-O-rhamnoside, 3-O-β-rutinoside (K-3-rh)/ 6-hydroxykaempferol 3, 6-di-O-β-D-glucoside, and kaempferide. Emerging evidence from modern pharmacological research suggests that kaempferol and its derivatives possess neuroprotective properties, particularly in the treatment of neurodegenerative diseases [28]. At the cellular level, kaempferol demonstrates anti-inflammatory effects. ...
... By obstructing the cation-selective channel (3a channel) that is present in cells infected with SARS-CoV-2, the innate yellow flavonoid kaempferol can hinder the outflow of K + , increase the level of Ca 2+ , and initiate the excretion of developed viruses (Imran et al. 2019). The NOD-like receptors (NLR) family pyrin domaincontaining 3 (NLRP3) inflammasome, the production of interleukin 1 (IL-1), and pyroptosis in lung cells are further processes that the 3a channel is thought to be involved in. ...
Article
Full-text available
The Coronavirus disease-2019 (COVID-19) pandemic is a global concern, with updated pharmacological therapeutic strategies needed. Cancer patients have been found to be more susceptible to severe COVID-19 and death, and COVID-19 can also lead to cancer progression. Traditional medicinal plants have long been used as anti-infection and anti-inflammatory agents, and Moringa oleifera (M. oleifera) is one such plant containing natural products such as kaempferol, quercetin, and hesperetin, which can reduce inflammatory responses and complications associated with viral infections and multiple cancers. This review article explores the cellular and molecular mechanisms of action of M. oleifera as an anti-COVID-19 and anti-inflammatory agent, and its potential role in reducing the risk of cancer progression in cancer patients with COVID-19. The article discusses the ability of M. oleifera to modulate NF-κB, MAPK, mTOR, NLRP3 inflammasome, and other inflammatory pathways, as well as the polyphenols and flavonoids like quercetin and kaempferol, that contribute to its anti-inflammatory properties. Overall, this review highlights the potential therapeutic benefits of M. oleifera in addressing COVID-19 and associated cancer progression. However, further investigations are necessary to fully understand the cellular and molecular mechanisms of action of M. oleifera and its natural products as anti-inflammatory, anti-COVID-19, and anti-cancer strategies.
... According to Ahmed et al. (2016), silver nanoparticles (AgNPs) include high-sensitivity bimolecular detection, diagnostic drug delivery, sanitization antimicrobials, therapeutics, catalysis, and microelectronics. Some important chemical compounds like phenolic acids (benzoic acid, gallic acid, caffeic acid), as well as flavonoids (kaempferol), are found in S. viarum, which exhibit the antimicrobial bioactivity (Imran et al. 2019;Asif and Khodadadi 2013). Hemashenpagam and Selvaraj (2009) reported higher antibacterial profile of S. viarum leaf extracts against Klebsiella pneumoniae, Pseudomonas aeruginosa, Staphylococcus aureus, Escherichia coli, and fungi such as Aspergillus niger, Candida albicans, and Penicillium sp. using the agar disc diffusion method. ...
... According to Ahmed et al. (2016), silver nanoparticles (AgNPs) include high-sensitivity bimolecular detection, diagnostic drug delivery, sanitization antimicrobials, therapeutics, catalysis, and microelectronics. Some important chemical compounds like phenolic acids (benzoic acid, gallic acid, caffeic acid), as well as flavonoids (kaempferol), are found in S. viarum, which exhibit the antimicrobial bioactivity (Imran et al. 2019;Asif and Khodadadi 2013). Hemashenpagam and Selvaraj (2009) reported higher antibacterial profile of S. viarum leaf extracts against Klebsiella pneumoniae, Pseudomonas aeruginosa, Staphylococcus aureus, Escherichia coli, and fungi such as Aspergillus niger, Candida albicans, and Penicillium sp. using the agar disc diffusion method. ...
Article
Full-text available
Solanum viarum, a perennial shrub, belongs to the family Solanaceae known for its therapeutic value worldwide. As a beneficial remedial plant, it is used for treating several disorders like dysentery, diabetes, inflammation, and respiratory disorders. Phytochemistry studies of this plant have shown the presence of steroidal glycoside alkaloids, including solasonine, solasodine, and solamargine. It also has flavonoids, saponins, minerals, and other substances. S. viarum extracts and compounds possess a variety of pharmacological effects, including antipyretic, antioxidant, antibacterial, insecticidal, analgesic, and anticancer activity. Most of the heavy metals accumulate in the aerial sections of the plant which is considered a potential phytoremediation, a highly effective method for the treatment of metal-polluted soils. We emphasize the forgoing outline of S. viarum, as well as its ethnomedicinal and ethnopharmacological applications, the chemistry of its secondary metabolites, and heavy metal toxicity. In addition to describing the antitumor activity of compounds and their mechanisms of action isolated from S. viarum, liabilities are also explained and illustrated, including any significant chemical or metabolic stability and toxicity risks. A comprehensive list of information was compiled from Science Direct, PubMed, Google Scholar, and Web of Science using different key phrases (traditional use, ethnomedicinal plants, western Himalaya, Himachal Pradesh, S viarum, and biological activity). According to the findings of this study, we hope that this review will inspire further studies along the drug discovery pathway of the chemicals extracted from the plant of S. viarum. Further, this review shows that ethnopharmacological information from ethnomedicinal plants can be a promising approach to drug discovery for cancer and diabetes. Graphical Abstract
Article
Zhujing pill (ZP) is a famous Chinese herbal formula that has been widely used to treat diabetic retinopathy, macular degeneration, retinitis pigmentosa and other fundus lesions. In this study, the material basis and mechanism of ZP in the treatment of fundus lesions were evaluated via the high‐performance liquid chromatography fingerprint, ultra‐performance liquid chromatography quadrupole time‐of‐flight mass spectrometry, network pharmacology and molecular docking. A total of 32 common components were found and 31 components were identified in 15 batches of ZP samples. Moreover, 134 common key targets and 17 putative active components that are connected to fundus lesions were identified. Molecular docking revealed that quercetin, kaempferol, isorhamnetin, 5‐ O ‐feruloylquinic acid, plantagoside and 2′‐acetylacteoside have the ability to interact with the core targets such as AKT1, TP53, TNF, IL‐6 and Jun. Our findings revealed that the therapeutic effects of ZP on fundus lesions are mediated by multiple components, targets and pathways, including at least six active ingredients and 11 targets. The study provides new ideas for further research on the material basis and mechanisms of traditional Chinese medicine prescriptions.
Article
Full-text available
The study aimed to determine the phenolic content and antioxidant capacity of five protein supplements of plant origin. The content and profile of phenolics were determined using the UHPLC-DAD-MS method, while antioxidant capacity (ABTS and DPPH assays) and total phenolic content (TPC) were evaluated using spectrophotometric tests. In the analyzed proteins, twenty-five polyphenols were detected, including eleven phenolic acids, thirteen flavonoids, and one ellagitannin. Hemp protein revealed the highest individual phenolics content and TPC value (1620 μg/g and 1.79 mg GAE/g, respectively). Also, hemp protein showed the highest antioxidant activity determined via ABTS (9.37 μmol TE/g) and DPPH (9.01 μmol TE/g) assays. The contents of p-coumaric acid, m-coumaric acid, kaempferol, rutin, isorhamnetin-3-O-rutinoside, kaempferol-3-O-rutinoside, and TPC value were significantly correlated with antioxidant activity assays. Our findings indicate that plant-based protein supplements are a valuable source of phenols and can also be used in research related to precision medicine, nutrigenetics, and nutrigenomics. This will benefit future health promotion and personalized nutrition in the prevention of chronic diseases.
Article
Full-text available
Chemotherapy-induced cognitive impairment or “chemobrain” is a prevalent long-term complication of chemotherapy and one of the more devastating. Most of the studies performed so far to identify the cognitive dysfunctions induced by antineoplastic chemotherapies have been focused on treatment with anthracyclines, frequently administered to breast cancer patients, a population that, after treatment, shows a high possibility of long survival and, consequently, of chemobrain development. In the last few years, different possible strategies have been explored to prevent or reduce chemobrain induced by the anthracycline doxorubicin (DOX), known to promote oxidative stress and inflammation, which have been strongly implicated in the development of this brain dysfunction. Here, we have critically analyzed the results of the preclinical studies from the last few years that have evaluated the potential of phenolic compounds (PheCs), a large class of natural products able to exert powerful antioxidant and anti-inflammatory activities, in inhibiting DOX-induced chemobrain. Several PheCs belonging to different classes have been shown to be able to revert DOX-induced brain morphological damages and deficits associated with learning, memory, and exploratory behavior. We have analyzed the biological and molecular mechanisms implicated and suggested possible future perspectives in this research area.
Article
In this study, we investigated the efficacy of kaempferol (a flavonoid found in plants and plant-derived foods such as kale, beans, tea, spinach and broccoli) on vascular contractibility and aimed to clarify the detailed mechanism underlying the relaxation. Isometric contractions of divested muscles were stored and linked with western blot analysis which was carried out to estimate the phosphorylation of myosin phosphatase targeting subunit 1 (MYPT1) and phosphorylation-dependent inhibitory protein for myosin phosphatase (CPI-17) and to estimate the effect of kaempferol on the RhoA/ROCK/CPI-17 pathway. Kaempferol conspicuously impeded phorbol ester-, fluoride- and a thromboxane mimetic-derived contractions regardless of endothelial nitric oxide synthesis, indicating its direct effect on smooth muscles. It also conspicuously impeded the fluoride-derived elevation in phospho-MYPT1 rather than phospho-CPI-17 levels and phorbol 12,13-dibutyrate-derived increase in phospho-CPI-17 and phospho-ERK1/2 levels, suggesting the depression of PKC and MEK activities and subsequent phosphorylation of CPI-17 and ERK1/2. Taken together, these outcomes suggest that kaempferol-derived relaxation incorporates myosin phosphatase retrieval and calcium desensitization, which appear to be modulated by CPI-17 dephosphorylation mainly through PKC inactivation.
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
Full-text available
Background: The epidemiological evidence for a relation between dietary polyphenol intake and depression is limited. Therefore, the aim of this study was to assess the association between habitual dietary intake of total polyphenols, their classes, subclasses and individual compounds and depressive symptoms among the participants of the Mediterranean healthy Eating, Lifestyle and Aging (MEAL) study. Methods: Demographic and dietary characteristics of 1572 adults living in southern Italy were analyzed. Food frequency questionnaires and Phenol-Explorer were used to calculate habitual dietary intakes of polyphenols. The Center for Epidemiologic Studies Depression Scale (CES-D-10) was used as screening tool for depressive symptoms. Multivariate logistic regression analyses were used to test associations and were expressed as odds ratio (OR) and 95% confidence intervals (CI). Results: A total of 509 individuals reported having depressive symptoms. Based on multivariate logistic regression analyses, total polyphenol intake was not associated with depressive symptoms. After adjustment for potential confounding factors, dietary intake of phenolic acid (OR = 0.64, 95% CI: 0.44, 0.93), flavanones (OR = 0.54, 95% CI: 0.32, 0.91), and anthocyanins (OR = 0.61, 95% CI: 0.42, 0.89) showed significant inverse association with depressive symptoms, when comparing the highest with the lowest quartile. Moreover, flavanones and anthocyanins, were associated with depressive symptoms in a dose-response manner. Among individual compounds, inverse association was observed for quercetin (OR = 0.53, 95% CI: 0.32, 0.86) and naringenin (OR = 0.51, 95% CI: 0.30, 0.85), for the highest versus lowest quartile of intake. When taking into consideration the major sources of the polyphenols, only citrus fruits and wine consumption was inversely associated with depressive symptoms (Q4 vs. Q1: OR= 0.51, 95% CI: 0.35, 0.75; Q4 vs. Q1: OR = 0.53, 95% CI: 0.38, 0.74, respectively). Conclusions: Higher dietary intake of flavonoid may be inversely associated with depressive symptoms. Further studies are needed to definitively confirm these observed associations.
Article
Full-text available
The main dietary sources of polyphenols are reviewed, and the daily intake is calculated for a given diet containing some common fruits, vegetables and beverages. Phenolic acids account for about one third of the total intake and flavonoids account for the remaining two thirds. The most abundant flavonoids in the diet are flavanols (catechins plus proanthocyanidins), anthocyanins and their oxidation products. The main polyphenol dietary sources are fruit and beverages (fruit juice, wine, tea, coffee, chocolate and beer) and, to a lesser extent vegetables, dry legumes and cereals. The total intake is ∼1 g/d. Large uncertainties remain due to the lack of comprehensive data on the content of some of the main polyphenol classes in food. Bioavailability studies in humans are discussed. The maximum concentration in plasma rarely exceeds 1 μM after the consumption of 10–100 mg of a single phenolic compound. However, the total plasma phenol concentration is probably higher due to the presence of metabolites formed in the body's tissues or by the colonic microflora. These metabolites are still largely unknown and not accounted for. Both chemical and biochemical factors that affect the absorption and metabolism of polyphenols are reviewed, with particular emphasis on flavonoid glycosides. A better understanding of these factors is essential to explain the large variations in bioavailability observed among polyphenols and among individuals.
Article
Full-text available
There is growing evidence that chronic hyperglycemia leads to the formation of advanced glycation end products (AGEs) which exerts its effect via interaction with the receptor for advanced glycation end products (RAGE). AGE-RAGE activation results in oxidative stress and inflammation. It is well known that this mechanism is involved in the pathogenesis of cardiovascular disease in diabetes. Kaempferol, a dietary flavonoid, is known to possess antioxidant, anti-apoptotic, and anti-inflammatory activities. However, little is known about the effect of kaempferol on myocardial ischemia-reperfusion (IR) injury in diabetic rats. Diabetes was induced in male albino Wistar rats using streptozotocin (70 mg/kg; i.p.), and rats with glucose level >250 mg/dL were considered as diabetic. Diabetic rats were treated with vehicle (2 mL/kg; i.p.) and kaempferol (20 mg/kg; i.p.) daily for a period of 28 days and on the 28th day, ischemia was produced by one-stage ligation of the left anterior descending coronary artery for 45 min followed by reperfusion for 60 min. After completion of surgery, rats were sacrificed and the heart tissue was processed for biochemical, morphological, and molecular studies. Kaempferol pretreatment significantly reduced hyperglycemia, maintained hemodynamic function, suppressed AGE-RAGE axis activation, normalized oxidative stress, and preserved morphological alterations. In addition, there was decreased level of inflammatory markers (tumor necrosis factor-α (TNF-α), interleukin-6 (IL-6), and NF-κB), inhibition of active c-Jun N-terminal kinase (JNK) and p38 proteins, and activation of Extracellular signal regulated kinase 1/2 (ERK1/2) a prosurvival kinase. Furthermore, it also attenuated apoptosis by reducing the expression of pro-apoptotic proteins (Bax and Caspase-3), Terminal deoxynucleotidyl transferase dUTP nick end labeling (TUNEL) positive cells, and increasing the level of anti-apoptotic protein (Bcl-2). In conclusion, kaempferol attenuated myocardial ischemia-reperfusion injury in diabetic rats by reducing AGE-RAGE/ mitogen activated protein kinase (MAPK) induced oxidative stress and inflammation.
Article
Full-text available
Background: Genomic DNA methylation plays an important role in both the occurrence and development of bladder cancer. Kaempferol (Kae), a natural flavonoid that is present in many fruits and vegetables, exhibits potent anti-cancer effects in bladder cancer. Similar to other flavonoids, Kae possesses a flavan nucleus in its structure. This structure was reported to inhibit DNA methylation by suppressing DNA methyltransferases (DNMTs). However, whether Kae can inhibit DNA methylation remains unclear. Methods: Nude mice bearing bladder cancer were treated with Kae for 31 days. The genomic DNA was extracted from xenografts and the methylation changes was determined using an Illumina Infinium HumanMethylation 450 BeadChip Array. The ubiquitination was detected using immuno-precipitation assay. Results: Our data indicated that Kae modulated DNA methylation in bladder cancer, inducing 103 differential DNA methylation positions (dDMPs) associated with genes (50 hyper-methylated and 53 hypo-methylated). DNA methylation is mostly relied on the levels of DNMTs. We observed that Kae specifically inhibited the protein levels of DNMT3B without altering the expression of DNMT1 or DNMT3A. However, Kae did not downregulate the transcription of DNMT3B. Interestingly, we observed that Kae induced a premature degradation of DNMT3B by inhibiting protein synthesis with cycloheximide (CHX). By blocking proteasome with MG132, we observed that Kae induced an increased ubiquitination of DNMT3B. These results suggested that Kae could induce the degradation of DNMT3B through ubiquitin-proteasome pathway. Conclusion: Our data indicated that Kae is a novel DNMT3B inhibitor, which may promote the degradation of DNMT3B in bladder cancer.
Article
Full-text available
Diabetic retinopathy (DR) is one of the common and specific microvascular complications of diabetes. This study aimed to investigate the anti-angiogenic effect of kaempferol and explore its underlying molecular mechanisms. The mRNA expression level of vascular endothelial growth factor (VEGF) and placenta growth factor (PGF) and the concentrations of secreted VEGF and PGF were measured by qTR-PCR and ELISA assay, respectively. Human retinal endothelial cells (HRECs) proliferation, migration, and sprouting were measured by CCK-8 and transwell, scratching wound, and tube formation assays, respectively. Protein levels were determined by western blot. High glucose (25 mM) increased the mRNA expression levels of VEGF and PGF as well as the concentrations of secreted VEGF and PGF in HRECs, which can be antagonized by kaempferol (25 µM). Kaempferol (5-25 µM) significantly suppressed cell proliferation, migration, migration distance and sprouting of HRECs under high glucose condition. The anti-angiogenic effect of kaempferol was mediated via downregulating the expression of PI3K and inhibiting the activation of Erk1/2, Src, and Akt1. This study indicates that kaempferol suppressed angiogenesis of HRECs via targeting VEGF and PGF to inhibit the activation of Src-Akt1-Erk1/2 signaling pathway. The results suggest that kaempferol may be a potential drug for better management of DR.
Article
Full-text available
Cervical cancer is one of the most frequent cancers in women worldwide. Defects in the apoptotic pathways are responsible for both the disease pathogenesis and its therapy resistance. It is thus a good candidate for treatment by pro-apoptotic agents. Kaempferol as a flavonoid has antioxidant and anti-tumor properties. Kaempferol has been shown to induce apoptosis and cell death in cancer cells. However, due to the problems in the treatment of cervical cancer, this study is designed to investigate the molecular mechanism by which kaempferol suppresses the growth of cervical cancer HeLa cell as compared with HFF cells (normal cells). Cells treated with kaempferol (12–100 μM) and 5-FU (1–10 μM), as the positive control, up to 72 h. Cell viability was determined by MTT assay and real time PCR was used to investigate apoptosis and telomerase genes expression. The results showed that kaempferol decreased cell viability as concentration- and time-dependently. IC50 values were 10.48 μM for HeLa and 707.00 μM for HFF cells, as compared with 1.40 μM and 16.38 μM for 5-FU after 72 h treatment, respectively. Also, kaempferol induced cellular apoptosis and aging through down-regulating the PI3K/AKT and hTERT pathways. This study suggests that kaempferol may be a useful adjuvant therapeutic agent in the treatment of cervical cancer.
Article
Kaempferol possesses a wide range of therapeutic properties, including antioxidant, anti-inflammatory, and anticancer properties. The present study sought to evaluate the effects and possible pharmacological mechanisms of kaempferol on interleukin (IL)-32-induced monocyte-macrophage differentiation. In this study, we performed flow cytometry assay, immunocytochemical staining, quantitative real-time PCR, enzyme-linked immuno sorbent assay, caspase-1 assay, and Western blotting to observe the effects and underlying mechanisms of kaempferol using the human monocyte cell line THP-1. The flow cytometry, immunocytochemical staining, and real-time PCR results show that kaempferol attenuated IL-32-induced monocyte differentiation to product macrophage-like cells. Kaempferol decreased the production and mRNA expression of pro-inflammatory cytokines, in this case thymic stromal lymphopoietin (TSLP), IL-1β, tumor necrosis factor (TNF)-α, and IL-8. Furthermore, kaempferol inhibited the IL-32-induced activation of p38 and nuclear factor-κB in a dose-dependent manner in THP-1 cells. Kaempferol also ameliorated the lipopolysaccharide-induced production of the inflammatory mediators TSLP, IL-1β, TNF-α, IL-8, and nitric oxide of macrophage-like cells differentiated by IL-32. In brief, our findings may provide new mechanistic insights into the anti-inflammatory effects of kaempferol.
Article
Recent evidence has suggested that flavonoid and lignan intake may be associated with decreased risk of chronic and degenerative diseases. The aim of this meta-analysis was to assess the association between dietary flavonoid and lignan intake and all-cause and cardiovascular disease (CVD) mortality in prospective cohort studies. A systematic search was conducted in electronic databases to identify studies published from January 1996 to December 2015 that satisfied inclusion/exclusion criteria. Risk ratios and 95% confidence intervals were extracted and analyzed using a random-effects model. Nonlinear dose-response analysis was modeled by using restricted cubic splines. The inclusion criteria were met by 22 prospective studies exploring various flavonoid and lignan classes. Compared with lower intake, high consumption of total flavonoids was associated with decreased risk of all-cause mortality (risk ratio = 0.74, 95% confidence intervals: 0.55, 0.99), while a 100-mg/day increment in intake led to a (linear) decreased risk of 6% and 4% of all-cause and CVD mortality, respectively. Among flavonoid classes, significant results were obtained for intakes of flavonols, flavones, flavanones, anthocyanidins, and proanthocyanidins. Only limited evidence was available on flavonoid classes and lignans and all-cause mortality. Findings from this meta-analysis indicated that dietary flavonoids are associated with decreased risk of all-cause and CVD mortality.
Article
Flavonoids are dietary intakes which are bestowed with several health benefits. The most studied property of flavonoids is their antioxidant efficacy. Among the chosen flavonoids Quercetin, Kaempferol and Myricetin is catagorized as flavonols whereas Apigenin and Luteolin belong to the flavone group. In the present study anti-cancer properties of flavonoids are investigated on the basis of their binding efficacy to ct-DNA and their ability to induce cytotoxicity in K562 leukaemic cells. The binding affinities of the flavonoids with calf thymus DNA (ct-DNA) are in the order Quercetin > Myricetin > Luteolin > Kaempferol > Apigenin. Quercetin with fewer OH than myricetin has higher affinity towards DNA suggesting that the number and position of OH influence the binding efficacies of flavonoids to ct-DNA. CD spectra and EtBr displacement studies evidence myricetin and apigenin to be stronger intercalators of DNA compared to quercetin. From comet assay results it is observed that quercetin and myricetin when used in combination induce higher DNA damage in K562 leukemic cells than when tested individually. Higher binding efficacy has been recorded for quercetin to DNA at lower pH, which is the micro environment of cancerous cells, and hence quercetin can act as a potential anti-cancer agent. Presence of Cu also increases cellular damage as recorded by comet assay.
Article
Kaempferol, a very common type of dietary flavonoids, has been found to exert antioxidative and anti-inflammatory properties. The purpose of our investigation was designed to reveal the effect of kaempferol on H9N2 influenza virus-induced inflammation in vivo and in vitro. In vivo, BALB/C mice were infected intranasally with H9N2 influenza virus with or without kaempferol treatment to induce acute lung injury (ALI) model. In vitro, MH-S cells were infected with H9N2 influenza virus with or without kaempferol treatment. In vivo, kaempferol treatment attenuated pulmonary edema, the W/D mass ratio, pulmonary capillary permeability, myeloperoxidase (MPO) activity, and the numbers of inflammatory cells. Kaempferol reduced ROS and Malondialdehyde (MDA) production, and increased the superoxide dismutase (SOD) activity. Kaempferol also reduced overproduction of TNF-α, IL-1β and IL-6. In addition, kaempferol decreased the H9N2 viral titre. In vitro, ROS, MDA, TNF-α, IL-1β and IL-6 was also reduced by kaempferol. Moreover, our data showed that kaempferol significantly inhibited the upregulation of toll-like receptor 4 (TLR4), myeloid differentiation factor 88 (MyD88), phosphorylation level of IκBα and nuclear factor-κB (NF-κB) p65, NF-κB p65 DNA binding activity, and phosphorylation level of MAPKs, both in vivo and in vitro. These results suggest that kaempferol exhibits a protective effect on H9N2 virus-induced inflammation via suppression of TLR4/MyD88-mediated NF-κB and MAPKs pathways, and kaempferol may be considered as an effective drug for the potential treatment of influenza virus-induced ALI.