Next Article in Journal
Crosstalk between Opioid and Anti-Opioid Systems: An Overview and Its Possible Therapeutic Significance
Previous Article in Journal
Clinical Translational Potential in Skin Wound Regeneration for Adipose-Derived, Blood-Derived, and Cellulose Materials: Cells, Exosomes, and Hydrogels
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Broad-Spectrum Preclinical Antitumor Activity of Chrysin: Current Trends and Future Perspectives

1
Department of Anatomical Sciences, School of Medicine, Shiraz University of Medical Sciences, Shiraz 7134814336, Iran
2
Cancer Science Institute of Singapore and Department of Pharmacology, Yong Loo Lin School of Medicine, National University of Singapore, Singapore 117599, Singapore
3
Department of Cell and Molecular Biology, Faculty of Biological Sciences, North Tehran Branch, IslamicAzad University, Tehran 165115331, Iran
4
Young Researchers and Elite Club, Tehran Medical Sciences, Islamic Azad University, Tehran 1916893813, Iran
5
Nursing and Midwifery Department, Islamic Azad University, Tehran Medical Sciences Branch, Tehran 1916893813, Iran
6
Sabanci University Nanotechnology Research and Application Center (SUNUM), Tuzla, Istanbul 34956, Turkey
7
Faculty of Veterinary Medicine, University of Tabriz, Tabriz 5166616471, Iran
*
Authors to whom correspondence should be addressed.
Submission received: 6 September 2020 / Revised: 21 September 2020 / Accepted: 23 September 2020 / Published: 27 September 2020

Abstract

:
Pharmacological profile of phytochemicals has attracted much attention to their use in disease therapy. Since cancer is a major problem for public health with high mortality and morbidity worldwide, experiments have focused on revealing the anti-tumor activity of natural products. Flavonoids comprise a large family of natural products with different categories. Chrysin is a hydroxylated flavonoid belonging to the flavone category. Chrysin has demonstrated great potential in treating different disorders, due to possessing biological and therapeutic activities, such as antioxidant, anti-inflammatory, hepatoprotective, neuroprotective, etc. Over recent years, the anti-tumor activity of chrysin has been investigated, and in the present review, we provide a mechanistic discussion of the inhibitory effect of chrysin on proliferation and invasion of different cancer cells. Molecular pathways, such as Notch1, microRNAs, signal transducer and activator of transcription 3 (STAT3), nuclear factor-kappaB (NF-κB), PI3K/Akt, MAPK, etc., as targets of chrysin are discussed. The efficiency of chrysin in promoting anti-tumor activity of chemotherapeutic agents and suppressing drug resistance is described. Moreover, poor bioavailability, as one of the drawbacks of chrysin, is improved using various nanocarriers, such as micelles, polymeric nanoparticles, etc. This updated review will provide a direction for further studies in evaluating the anti-tumor activity of chrysin.

1. Introduction

Average living standards and access to sufficient healthcare have led to an increase in life expectancy in most regions of the world [1,2]. Although communicable disease-related deaths have been reduced as a result of medical improvements, we have witnessed a 40% increase in cancer-related deaths in recent years. It seems that the number of patients with cancer will increase in the future, and there will be up to 13 million cancer-related deaths by 2030. There are different problems in providing effective cancer therapy, such as the insufficiency of currently applied treatments, lack of early diagnosis, and poor understanding of signaling networks involved in cancer malignancy. In spite of significant attempts in knowing factors contributing to cancer progression, there is not still an effective treatment for cancer [3,4,5]. This is due to the fact that each cancer type has its own features; for instance, cancer cells are different in terms of proliferation, metastasis, and dependence on molecular pathways. Furthermore, cancer cells can obtain resistance to currently applied chemotherapeutic agents [6,7]. Therefore, a novel agent capable of suppressing cancer growth and metastasis and preventing drug resistance is important. In the present review, we aim to reveal the anti-tumor activity of chrysin, as a naturally occurring compound against different cancers. We discuss the various molecular pathways that are affected by chrysin in cancer to direct further studies for investigating more signaling networks. In addition, we describe the role of chrysin in overcoming drug resistance in cancer therapy, which is a major problem in the clinic. Finally, we provide strategies in promoting the anti-tumor activity of chrysin using nanoparticles to enhance bioavailability and therapeutic effects of chrysin.

2. Role of Natural Products in Cancer Therapy

Nature is a rich source of compounds with different pharmacological activities [8,9,10,11,12]. The special view towards nature is due to the presence of anti-tumor agents with low toxicity, and capable of suppressing a wide variety of cancers [13,14,15,16,17,18]. Furthermore, natural products are more affordable compared to synthetic drugs. It seems that newly introduced anti-tumor drugs have high similarity to natural anti-tumor compounds. Therefore, identifying novel phytochemicals, making changes in their structure to promote their therapeutic effect, and introducing into the market can be considered as a new way in effective cancer therapy. Newly published experiments have clearly demonstrated the potential of phytochemicals in cancer therapy. The proliferation of cancer cells is suppressed upon the administration of natural anti-tumor compounds [19,20]. Apoptosis and cell cycle arrest can be induced via p53 up-regulation [21]. Based on the fact that poor bioavailability is one of the drawbacks of natural products, using nanoscale delivery systems can exponentially promote their anti-tumor activity against cancer cells for both in vitro and in vivo experiments [22,23,24]. In cancer cells, checkpoint gene expression enhances that provides uncontrolled growth. It has been reported that the administration of natural products is correlated with a decrease in checkpoint expression, and subsequent decrease in proliferation of cancer cells [25]. DNA damage, as well as the activation of both intrinsic and extrinsic pathways of apoptosis, occur during natural product administration in cancer therapy [26]. It is worth mentioning that naturally occurring compounds can promote the efficiency of chemotherapeutic agents in cancer therapy [27,28,29]. For instance, quercetin sensitizes prostate cancer cells to paclitaxel chemotherapy by enhancing reactive oxygen species (ROS) production, stimulation of endoplasmic reticulum (ER) stress, and activation of apoptosis [30]. Molecular pathways. such as MAPK and JNK, are regulated by natural products in apoptosis induction [31]. In addition to proliferation, migration, and invasion of cancer cells can be negatively targeted by natural products [32,33,34]. Increasing evidence confirms the role of epithelial-to-mesenchymal transition (EMT) in cancer metastasis [35,36,37]. Natural products are capable of suppressing the migration of cancer cells by EMT inhibition via down-regulation of upstream molecular pathways, such as Snail and STAT3 [38,39,40].
Taking everything into account, studies agree with the fact that natural products are versatile compounds in cancer therapy, and due to their capacity in targeting various molecular pathways in cancer therapy [41,42,43,44,45], they can be considered as potential agents in the field of cancer therapy. In the next sections, we focus on chrysin as an efficient anti-tumor agent in different cancers.

3. Chrysin: An Overview of Chemistry, Sources, and Pharmacokinetics

Flavonoids are the largest group of plant secondary metabolites with favorable health-promoting effects [46,47,48,49]. The interest in flavonoids has been increased, since these valuable compounds act through various physiological mechanisms and affect a wide variety of signaling networks. Dietary intake of flavonoids is estimated to be 50 and 800 mg per day [50,51]. Chrysin is a hydroxylated flavonoid belonging to flavone class, and is extensively found in sources, such as honey, propolis, and plant species [52,53]. Noteworthy, chrysin occurs in natural sources with different concentrations. For instance, the concentration of chrysin in honeydew honey is 0.10 mg/kg, while it has a higher concentration (5.3 mg/kg) in forest honeys [54]. The content of chrysin in propolis is estimated to be 25 g/L [55]. Chrysin concentration in mushrooms is at the range of 0.17–0.34 mg/kg [56]. The IUPAC name of chrysin is 5,7-dihydroxy-2-phenyl-4H-chromen4-one and 5,7-dihydroxyflavone. Figure 1 demonstrates the chemical structure of chrysin. The chrysin structure has similarities and differences with the flavonoid family. Structurally, chrysin has two benzene rings (A and B) with one oxygen consisting of a heterocyclic ring. Chrysin lacks a 3-carbon hydroxyl group, but it has 2–3 double-bound carbon with a carbonyl group attached to 4th carbon. The chemical structure of chrysin demonstrates that it has –OH group at 5th and 7th carbon atoms. There is a difference in the structure of chrysin and other flavonoids, so that chrysin does not possess any oxygenation in ring B (Figure 1). It has been reported that changes in ring A of chrysin account for the generation of different derivatives of chrysin, such as wogonin, baicalein, and oroxylin [57].
Accumulating data demonstrates that poor absorption, rapid metabolism, and systemic elimination are responsible for poor bioavailability of chrysin in humans that, subsequently, restrict its therapeutic effects [58]. It is worth mentioning that oxidation in intestinal and hepatic cells is not responsible for the metabolism of chrysin in the body. In contrast, conjugation pathways, such as glucuronidation and sulfation catalyze chrysin. Enzymes, such as P-PST, M-PST, and UGT1A6, contribute to the metabolism of chrysin, and their high affinity for chrysin can justify the poor bioavailability of this natural compound. Clinical studies have shown that the plasma concentration of chrysin following oral administration is very low [59]. Notably, serum concentrations of chrysin have not been reported yet, but it can be predicted based on other flavonoids. Since flavonoid aglycones demonstrate serum concentration as low as 1 µmol/L [60], the serum concentration of chrysin would be at the range of nanomolar. The studies related to the absorption of chrysin demonstrate that its sulfation and glucuronidation limit the absorption of this valuable compound in the intestine. MRP2 transporters are involved in the efflux of chrysin metabolites from the intestine, and in the lumen, sulfatases and glucuronidases hydrolyze metabolites into chrysin. This leads to the emergence of chrysin in stool, but high contents of chrysin in stool demonstrates that it has low absorption [61]. Some strategies have been applied in promoting bioavailability and absorption of chrysin, such as using nanoscale delivery systems [62].

4. Chrysin and Its Pharmacological Activities

In previous sections, we provided explanations about the role of natural products in cancer therapy, and then, we introduced the chemistry and pharmacokinetics of chrysin. In this section, we aim to describe the pharmacological activities of chrysin, based on the newly published article—which is summarized in Table 1.
Increasing evidence demonstrates that chrysin possesses health-promoting effects, including antioxidant [63,64], anti-inflammatory [65], anti-diabetes [66], neuroprotective [67], hepatoprotective [68], cardioprotective [69], lipid-lowering effect [70], etc. These therapeutic effects have made chrysin as a suitable option in disease therapy. Non-alcoholic fatty liver disease (NAFLD) is one of the most common metabolic disorders, and to date, natural products have shown great potential in the alleviation of NAFLD. Similarly, a recently recorded article has revealed that chrysin administration (25, 50, and 100 mg/kg) alleviates NAFLD in rats via reducing serum fasting glucose that subsequently improves insulin resistance and dyslipidemia. Noteworthy, chrysin can significantly diminish liver weight by reducing hepatic free fatty acids, triglyceride, and cholesterol content. Anti-inflammatory and antioxidant activities of chrysin are also involved in the amelioration of NAFLD via decreasing lobular inflammation, steatosis, and carbonyl content [71]. Many reports demonstrate that chrysin can be beneficial in reducing acetaminophen-mediated hepatotoxicity in rats. In this regard, chrysin reduces levels of pro-inflammatory cytokines, such as tumor necrosis factor-α (TNF-α) and interleukin-2 (IL-2). The ameliorative effect of chrysin on acetaminophen-mediated hepatotoxicity seems to be dose-dependent with more therapeutic effects at higher concentrations [72]. In addition to hepatoprotective activity, chrysin has shown potential neuroprotective effects. One of the complications causing neuronal cell death is ischemic-reperfusion (I/R) injury. Inflammation and oxidative stress are two main mechanisms involved in I/R injury [73,74,75]. Chrysin administration (10 and 20 mg/kg) reduces pro-inflammatory factors (TNF-α, IL-1β, and IL-6) and oxidative stress to alleviate cerebral I/R injury. Investigation of molecular pathways reveals that the induction of the PI3K/Akt signaling pathway by chrysin contributes to a reduction in oxidative stress and inflammation during cerebral I/R injury [76]. The inhibitory effect of chrysin on inflammation and oxidative stress is also important in Parkinson’s disease (PD) treatment [77]. Chrysin (25, 50, and 100 mg/kg) improves cognitive capacity, inflammation, and apoptosis to ameliorate traumatic brain injury (TBI) [78]. Overall, the literature confirms the health-promoting and therapeutic effects of chrysin that are important in disease therapy, and the effect of this valuable compound on molecular pathways (Figure 2) [79,80,81,82]. In the next sections, we specifically discuss the role of chrysin in cancer therapy [83,84].

5. Chrysin and Cancer

5.1. Breast Cancer

Breast cancer is the most common and malignant cancer in women [94,95,96]. Recurrence and chemoresistance have restricted the efficacy of currently applied treatment in breast cancer therapy [97,98,99,100]. Natural products have demonstrated an excellent inhibitory effect on both proliferation and metastasis of breast cancer [101,102,103,104]. A combination of chrysin and silibinin is beneficial in suppressing breast cancer malignancy via decreasing cancer proliferation. Furthermore, chrysin and silibinin induced cell cycle arrest via down-regulation of cyclin D1 and hTERT [105]. The epidermal growth factor receptor (EGFR) is considered as a potential target in cancer therapy [106]. Standard chemotherapy reduces the replication of cancer cells, but EGFR inhibitors are capable of cancer proliferation and survival [107]. Therefore, using EGFR inhibitors, such as antibody-based immunoconjugates, monoclonal antibodies, antisense oligonucleotides, and small molecules is preferred to chemotherapy [108]. A new derivative of chrysin known as CHM-04 has been synthesized with affinity to EGFR. It seems that CHM-04 is a potent inhibitor of EGFR with more efficiency compared to chemotherapeutic agents in suppressing cancer malignancy. In triple-negative breast cancer cells treated with chrysin, sphere formation ability, proliferation, and migration are substantially suppressed that can be attributed to the inhibitory effect of CHM-04 on EGFR [109].
Low oxygen level is known as hypoxia, and is a common feature of solid tumors. Increasing evidence demonstrates that hypoxia is responsible for the growth and progression of cancer cells, and it is one of the best targets in cancer therapy [110,111,112]. Noteworthy, clinical studies revealed the relationship between hypoxia and cancer progression and metastasis [113,114]. In hypoxia, vascular endothelial growth factor (VEGF) is induced that promotes proliferation and invasion of cancer cells. Furthermore, hypoxia adaptation is mediated by hypoxia-inducible factor-1 (HIF-1) that is an efficient target in cancer therapy. In addition to HIF-1, other molecular pathways, such as signal transducer and activator of transcription 3 (STAT3), play a key role in hypoxia-mediated VEGF gene expression [115,116,117,118]. Administration of chrysin is associated with the disruption of hypoxia-induced VEGF gene expression. Moreover, chrysin is capable of reducing STAT3 phosphorylation in hypoxic conditions without affecting the HIF-1α protein level. In vitro and in vivo experiments agree with the fact that chrysin is a potent agent in suppressing metastasis and proliferation of breast cancer cells during hypoxic conditions, since chrysin abrogated lung metastasis of breast cancer cells [119].
Increasing evidence demonstrates that combination therapy is of interest in promoting the anti-tumor activity of agents. Although chrysin has demonstrated great potential in suppressing proliferation and metastasis of cancer cells, its anti-tumor activity can be promoted by combination therapy. Metformin, as an anti-diabetic agent, has been applied in cancer therapy, due to its capacity in inhibiting proliferation, metastasis, and induction of apoptosis, and cell cycle arrest [120,121]. It seems that combination therapy of breast cancer cells using chrysin and metformin exerts a synergistic effect and is more efficient compared to chrysin alone. Cyclin D1 and hTERT are down-regulated by chrysin and metformin in breast cancer therapy [122].

5.2. Lung Cancer

International Agency for Research on Cancer has considered nickel as one of the carcinogenic agents [123,124,125]. Exposing to nickel-containing compounds is correlated with the risk of lung cancer development [126,127]. Enhancing ROS levels, inflammation induction, epigenetic gene regulation, and stimulation of signaling pathways are positively affected by nickel in cancer development [128,129,130]. Furthermore, activation of toll-like receptors (TLRs) is associated with cancer development [131,132]. The nuclear factor-kappaB (NF-κB) signaling pathway promotes inflammation and cancer progression [133]. A report has evaluated and compared the efficiency of five natural products, including quercetin, chrysin, curcumin, apigenin, and luteolin. Among them, quercetin and chrysin demonstrated the highest efficacy in lung cancer treatment. A combination of quercetin and chrysin reduced levels of pro-inflammatory factors, such as IL-1β, Il-6, TNF-α, and IL-10, via NF-κB down-regulation. Furthermore, chrysin and quercetin decreased expressions of Myd88 and TLR4, as well as MMP-9, to suppress the viability and metastasis of lung cancer cells [134].

5.3. Prostate Cancer

Prostate cancer (PC) is one of the most common cancers in men that is responsible for 21% of cancer cases and 8% of cancer-related deaths in the United States [135,136,137]. Chemotherapy, radiotherapy, and prostatectomy are strategies in PC therapy, but recurrence and resistance of PC cells are problems, requiring novel strategies in PC therapy [138,139]. Increasing evidence demonstrates that PI3K/Akt and MAPK signaling pathways account for an increase in proliferation and metastasis of cancer cells, and their inhibition is important in cancer therapy [140,141,142,143]. In PC cells, chrysin down-regulates the expression of the PI3K/Akt pathway to interrupt the proliferation of PC cells. Furthermore, MAPK down-regulation by chrysin leads to a decrease in PC proliferation. Chrysin is able to induce apoptosis in PC cells via mitochondrial dysfunction, so that after chrysin administration, an increase occurs in levels of ROS that, subsequently, impairs the integrity of the mitochondrial membrane, leading to cytochrome C release and apoptosis induction [144]. Noteworthy, in addition to mitochondria, ER can also participate in apoptosis. The primary role of ER is to preserve cell homeostasis and ensuring the correct conformation of proteins. ER stress occurs when levels of unfolded proteins exceed from the capacity of ER. This leads to the activation of unfolded protein response (UPR) that, subsequently, stimulates PRKR-like ER kinase (PERK), eukaryotic translation initiation factor 2α (eIF2α), and 78 kDa glucose-regulated protein (GRP78) [145,146,147,148]. Chrysin administration also impairs ER homeostasis to induce ER-mediated apoptosis in PC cells [144].

5.4. Ovarian Cancer

Ovarian cancer (OC) is the fifth leading cause of death in women, and is considered one of the most lethal gynecologic cancers [135,149]. Based on the experiments performed in the field of OC treatment, it seems that phytochemicals are potential therapeutic agents in this case [150,151]. In the previous section, we discussed that mitochondrial dysfunction leads to apoptosis induction [152]. Upon chrysin administration, an increase occurs in levels of ROS and cytoplasmic Ca2+ that mediate apoptosis induction in OC cells [153]. However, this study provides controversial results about the role of molecular pathways that needs to be explored in further studies. Accumulating data demonstrates that the PI3K/Akt signaling pathway contributes to cancer proliferation and metastasis. PI3K/Akt inhibition has been suggested in different experiments as a promising strategy in cancer therapy [154,155,156]. However, a previous study has shown that chrysin suppresses OC malignancy via PI3K/Akt and MAPK induction [153]. Therefore, further studies are required to shed some light on this area.

5.5. Gastric Cancer

Gastric cancer (GC) is the third leading cause of cancer death, with 783,000 deaths in 2018 [157,158,159,160]. Different factors are involved in GC progression, and ten-eleven translocation (TET) enzyme is one of them. TET enzymes contribute to the oxidation of 5-methylcytosine (5mC) to 5-hydroxymethylcytosine (5hmC) and participate in epigenetic modification [161]. Studies show the role of TET enzymes in GC development. For instance, TET1-mediated demethylation stimulates the aggressive behavior of GC cells [162]. TET2 exerts RASSF1A methylation to affect malignant cell activity [163]. Furthermore, down-regulation of TET3 has been shown in GC [164]. The effect of chrysin on GC cells has been investigated in vitro and in vivo. In MKN45 cells, chrysin promotes the expression of TET1 and 5hmC to stimulate apoptosis and disrupt migration and invasion of GC cells. Furthermore, TET1 deletion by CRISPR/Cas9 system in a mouse model leads to the development of GC, and chrysin administration can be considered as a promising strategy in GC suppression [165].
One of the properties of phytochemicals is their capability to regulate microRNA (miR) expression [166,167]. Briefly, miRs are non-coding parts of the genome that are not transcribed into protein [168]. Cellular mechanisms, such as proliferation, migration, differentiation, etc., are tightly regulated by miRs [169]. Disturbance in miR expression leads to the emergence of pathological conditions, particularly cancer [170,171]. Chrysin is capable of promoting the expression of miR-9 and Let-7a as onco-suppressor factors in cancer to inhibit the proliferation of GC cells. Using nanoparticles can significantly promote the ability of chrysin in enhancing miR-9 expression [172].

5.6. Cervical Cancer

Cervical cancer is one of the most common malignancies diagnosed in women [173,174,175]. Chronic infection with high-risk human papillomavirus and inherited polymorphism of cytokine genes are involved in cervical cancer emergence [176,177,178,179,180]. Hence, enhanced levels of cytokines participate in cervical cancer progression. Furthermore, EMT-related metastasis provides a poor prognosis of patients with cervical cancer [181]. Hence, anti-tumor compounds with a modulatory effect on inflammation can be beneficial in suppressing cervical cancer metastasis. Exposing cervical cancer cells into transforming growth factor-beta (TGF-β) is associated with enhanced levels of TNF-α, inflammation, and metastasis. As a consequence of inflammation, NF-κB is activated that induces Twist/EMT axis in cervical cancer metastasis. Chrysin (5, 10 and 20 µM) suppresses the aggressive behavior of cervical cancer cells in a dose-dependent manner. Down-regulation of NF-κB, and subsequent decrease in Twist/EMT are mediated by chrysin administration, negatively affecting cervical cancer metastasis [182].
Scutellaria discolor Colebr is a well-known medicinal plant species with therapeutic effects in treating different diseases [183]. There have been efforts in revealing bioactive compounds in this plant that are responsible for its pharmacological activities, particularly cancer. It has been reported that chrysin is the major bioactive component of this plant that provides the anti-tumor activity against cervical cancer cells. Induction of cell cycle arrest and apoptosis via up-regulation of caspase-3, caspase-9, and Bax are mediated by chrysin. Moreover, chrysin impairs the proper function of mitochondria via providing mitochondrial membrane depolarization, leading to reduced viability of cervical cancer cells [184].

5.7. Liver Cancer

Studies are in line with the fact that cancer cells are different from normal cells in terms of metabolism [185]. Aerobic glycolysis, or the Warburg effect, is one of the hallmarks of cancer that was first recognized in 1920 by Otto Heinrich Warburg [186]. In this process, regardless of oxygen levels, glucose is converted into lactate to meet the needs of cancer cells into energy, leading to their uncontrolled proliferation [187]. Different factors have been recognized to participate in changing the metabolism of cancer cells from the Krebs cycle to glycolysis, and hexokinases (HKs) are one of them [188,189]. A large body of evidence shows the important role of HK-2 in the Warburg effect in different cancers [190,191,192,193]. Chrysin administration (15, 30, and 60 mM) reduces the expression of HK-2 in hepatocellular carcinoma (HCC) cells to impair glucose uptake and lactate production. In addition to glycolysis metabolism impairment, the inhibitory effect of chrysin on HK-2 leads to apoptosis, so that chrysin disrupts the interaction of HK-2 and VDAC-1 on mitochondria that releases Bax from mitochondrial into the cytoplasm, leading to apoptosis induction. Notably, tumor xenografts treated with chrysin demonstrated a decrease in HK-2 levels in tissues [194].
The main pathway that is followed by chrysin in suppressing liver cancer survival is apoptosis induction. In this way, chrysin substantially enhances levels of ROS that, subsequently, disturbs mitochondrial function. Disruption in the integrity of the mitochondrial membrane leads to cytochrome C release into the cytoplasm, resulting in apoptotic cell death [195].
Increasing evidence is in agreement with the fact that the STAT3 signaling pathway participates in the proliferation and invasion of HCC cells [196,197,198,199]. Inhibition of STAT3 by anti-cancer agents is important in effective HCC therapy [200,201,202]. In HCC cells exposed to chrysin, a decrease occurs in sphere formation capacity. Investigation of molecular pathways reveals that STAT3 undergoes down-regulation upon chrysin administration. Notably, an upstream modulator of STAT3 known as SHP-1 is up-regulated by chrysin, and consequently, it decreases expression of STAT3, leading to inhibited sphere formation [203].

5.8. Melanoma

Melanoma is a highly resistant and malignant tumor of the skin that is responsible for about 3% of all cancer cases. Over the past decades, we have witnessed an increase in the occurrence of melanoma. Although melanoma accounts for 4% of all skin cancer cases, its aggressiveness and malignancy have led to comprising 80% of all deaths from skin cancer [204]. Melanoma, at the first stages, can be treated with surgery, but in an advanced stage, it metastasizes into other sites, making its treatment more complex [205,206,207]. Plant derived-natural compounds can be considered as potential agents in melanoma therapy, due to their ability in apoptosis and cell cycle induction, and inhibiting migration [208,209,210]. Chrysin is a potent agent in melanoma therapy, and this ability has been approved in vitro and in vivo. Chrysin stimulates apoptosis and cell cycle arrest (G2/M phase) in a dose-dependent manner. In tumor xenografts, chrysin decreases tumor growth by 60% after 14 days of treatment, while this number enhances to 70% after 21 days of treatment. Noteworthy, in melanoma therapy, chrysin promotes cytotoxicity activity of natural killer cells, macrophages, and cytotoxic T cells [211].
MMPs are involved in enhancing the invasion of cancer cells via extracellular matrix (ECM) degradation [212,213]. MMP-2 and MMP-9 provide metastasis of cancer cells into distant organs via degrading matrix collagen and basement membrane [214,215]. Chrysin (5–15 µM) suppresses metastasis of melanoma cells via down-regulation of MMP-2. Furthermore, N-cadherin and E-cadherin are respectively down-regulated and up-regulated upon chrysin administration in inhibiting melanoma invasion [182]. In previous sections, we discussed the oncogene role of NF-κB and PI3K/Akt signaling pathways in cancer. Chrysin treatment is associated with a decrease in expression of NF-κB and PI3K/Akt to suppress melanoma proliferation [182].

5.9. Bladder Cancer

The second most common type of tract cancer in developed countries is bladder cancer. Its incidence rate is around 400,000 cases, with approximately 160,000 death annually [135]. Chemotherapy is not suggested in bladder cancer therapy, due to side effects and chemoresistance [216]. Novel strategies can be developed for promoting the efficacy of chemotherapy in bladder cancer therapy, such as using phytochemicals with anti-tumor activity [217,218]. On the other hand, molecular pathways, such as STAT3 participate in bladder cancer progression [219]. STAT3 can individually promote the proliferation of bladder cancer cells [220], or it may be targeted by upstream mediators, such as Akt/ERK [221]. Administration of chrysin is correlated with an increase in ROS levels to down-regulate STAT3 expression. Furthermore, chrysin activates the intrinsic pathway of apoptosis via caspase-3 and caspase-9 up-regulation. Anti-apoptotic factors, such as Bcl-2, Mcl-1, and Bcl-xl undergo down-regulation by chrysin in bladder cancer cells. Notably, chrysin substantially diminishes survival by ER stress induction via stimulating UPR, PERK, ATF4, and elF2α [222].

5.10. Colorectal Cancer

Colorectal cancer (CRC) is a heterogeneous disease with a rise in the incidence rate in recent years. Both molecular and pathological properties determine the prognosis and response of CRC cells into therapy [223,224]. 5-fluorouracil (5-FU) is extensively applied in treating patients with CRC, but drug resistance and side effects have restricted its use [225,226]. Recently, chrysin has been considered as a substitution for 5-FU in CRC therapy. Chrysin administration (5–50 µM) is associated with a significant decrease in the viability of CRC cells [227]. An investigation into the molecular mechanisms demonstrates that autophagy is affected by chrysin in CRC therapy. Autophagy is a “self-digestion” process with stimulation upon stressful conditions, such as ER stress, mitochondrial damage, starvation, etc. [228,229]. Autophagic cell death is important in reducing the viability of cancer cells [230,231]. Chrysin enhances levels of light chain-3 II (LC-3II) to induce autophagy. Furthermore, by promoting ROS generation, chrysin down-regulates the expression of the mammalian target of rapamycin (mTOR) to stimulate autophagy, leading to a decrease in the viability of CRC cells [227].
It is worth mentioning that irradiation can improve the anti-tumor activity of chrysin against colon cancer cells. Irradiation technology is able to promote biological properties or physical features of biomolecules through structural modification [232,233,234]. Recently, chrysin and gamma irradiation have been co-applied in colon cancer therapy. Irradiation substantially enhances the cytotoxic activity of chrysin. This inhibitory effect against colon cancer cells is exerted via promoting ROS generation, inducing mitochondrial dysfunction, activation of a caspase cascade (caspase-3 and caspase-9), and stimulating cleavage of poly (adenosine diphosphate-ribose) polymerase (PARP) [235].
Peroxisome proliferator-activated receptor alpha (PPARα) is a crucial member of the superfamily of nuclear hormone receptors with regulatory effects on migration, proliferation, metabolism, etc. [236,237,238,239]. Increasing evidence demonstrates that using a specific ligand for stimulation of PARPα is of interest in suppressing cancer growth [240,241]. On the other hand, cytochrome P450 (CYPs) enzymes contribute to drug metabolism and are found in different organs of the body, such as lung, liver, etc. [242,243]. PARPα is able to regulate gene expression of CYPs, such as CYP3A4 and CYP2C8 [244]. Chrysin administration significantly enhances the expression of PARPα in cancer cells. This leads to a significant reduction in expression of CYP2S1 and CYP1B1, leading to decreased proliferation (cell cycle arrest) and migration of cancer cells [245].
A schematic summary on anti-tumor effects of chrysin in cancer is shown in Figure 3. Table 2 list chrysin administration in treating various cancers.

6. Chrysin, Chemotherapy and Drug Resistance

Chemotherapy is an inevitable part of cancer therapy, but its potential has been restricted in recent years, due to the resistance of cancer cells [259,260]. In fact, chemoresistance of cancer cells has urged scientists to seek new anti-tumor agents [261]. Based on the role of natural products in cancer treatment, they can be beneficial in sensitizing cancer cells into chemotherapy [262,263]. That is why these valuable agents have been extensively co-administered with chemotherapeutic agents in cancer therapy. Anti-tumor phytochemicals can suppress proliferation, metastasis, and malignant behavior of cancer cells that are in favor of chemotherapeutic agents [102,264,265]. In this section, we provide a discussion about the role of chrysin as a naturally occurring compound in reversing drug resistance.
Cisplatin is a well-known chemotherapeutic agent with clinical application. However, resistance is the most important reason for treatment failure with this agent in the clinic [266,267]. Various molecular pathways have been suggested to participate in cisplatin resistance, such as CLEC4M, miRs, lncRNAs, etc. [268,269]. In respect to the high anti-tumor activity of chrysin, this plant derived-natural compound can be advantageous in suppressing chemoresistance. Noteworthy, it has been reported that selenium-containing chrysin and quercetin derivatives are potent agents in reversing cisplatin resistance [270].
Docetaxel (DTX) is a commercially applied chemotherapeutic agent in treating lung cancer, breast cancer, gastric cancer, etc. DTX stimulates apoptosis and cell cycle arrest via attaching β-tubulin into microtubules and disrupting cancer growth [271]. Similar to other chemotherapeutic agents, cancer cells are capable of obtaining resistance to DTX [272]. Moreover, the anti-tumor activity of DTX can be improved by combinational therapy [273]. A combination of chrysin (20–100 µM) and DTX is advantageous in suppressing the proliferation of cancer cells, and inducing growth delay in tumor xenografts [274]. This is distributed to apoptosis induction by chrysin that, subsequently, sensitizes cancer cells into DTX chemotherapy [274].
P53 is a key player in apoptosis induction. It stimulates apoptosis in both transcription-dependent and transcription-independent manners. In the transcription-dependent pathway, down-stream genes of p53 are regulated to induce apoptosis in cancer cells [275,276,277]. Furthermore, p53 is capable of moving out of the nucleus, and interacting with mitochondria and its proteins, such as Bcl-2 and Bcl-xl, in apoptosis induction [278]. In liver cancer cells exposed to chrysin and cisplatin, an increase occurs in phosphorylation and accumulation of p53 via ERK1/2 up-regulation. Consequently, apoptotic factors, such as Bax and DR5, undergo up-regulation, while a decrease occurs in the expression of anti-apoptotic factor Bcl-2. The intrinsic pathway of apoptosis is activated via caspase-8 activation. Chrysin and cisplatin also induce the extrinsic pathway of apoptosis via releasing cytochrome C into the cytoplasm and activating caspase-9 [279].
Nuclear factor erythroid 2-related factor 2 (Nrf2) is an important signaling pathway involved in antioxidant activity against oxidative stress and other kinds of stresses [280,281,282]. Recently, much attention has been directed towards the role of Nrf2 in the chemoresistance of cancer cells [283]. Nrf2 follows different routes in exerting chemoresistance, such as enhancing expression of CD99 [284], inhibiting DNA damage [285], and reducing oxidative stress-mediated damage [286]. Therefore, Nrf2 inhibition is important in reducing chemoresistance. Chrysin administration (10 and 20 mM) promotes the sensitivity of cancer cells into doxorubicin chemotherapy. Further analysis reveals that Nrf2 undergoes down-regulation by chrysin in cancer cells. Furthermore, in reducing Nrf2 expression, chrysin down-regulates the expression of ERK and PI3K/Akt pathways—leading to an increase in the efficiency of doxorubicin in chemotherapy [287].

7. Chrysin-Loaded Nanoparticles in Cancer Therapy

Micelles have attracted much attention in cancer therapy, due to their potential to deliver anti-tumor agents [288,289]. Self-assembled micelles are amphiphilic copolymers with size at the range of 10–100 nm. Micelles have high cellular uptake and passive targeting functions to tumor known as enhanced permeability [290,291]. Recently, chrysin- and docetaxel-loaded micelles have been applied in enhancing the efficacy of chemotherapy. This co-delivery by micelles exerts a synergistic effect on chemotherapy and effectively suppresses migration and invasion of cancer stem cells. Chrysin- and docetaxel-loaded micelles enhance levels of ROS to impair cancer stem cell viability. Notably, enhanced the anti-tumor activity of chrysin and docetaxel against cancer cells is due to their enhanced accumulation in cancer cells by micelles [292]. Polymeric micelles have also been designed in co-delivery of chrysin and methotrexate in the chemotherapy of breast cancer cells. The idea of using a chemotherapeutic agent with a natural anti-tumor agent is that this combination is important in sensitizing cancer cells into chemotherapy. Using nanoparticles promotes cytotoxicity against cancer cells via enhancing cellular uptake. Based on the small size of polymeric micelles (around 55 nm), they can escape from macrophages and kidney filtration to reach into the tumor site, providing targeted delivery of anti-tumor compounds [293].
Another study has applied polyurea dendrimers for delivery of chrysin in ovarian cancer therapy. Polyurea dendrimers are three-dimensional polymers with urea moieties in the backbone and peripheral amine groups. They possess various beneficial properties, including water-solubility, biocompatibility, biodegradability, and pH-sensitivity, making them suitable options in drug delivery [294]. Furthermore, as cancer cells overexpress folate receptors on their surface [295,296], surface functionalization of nanoparticles with folate can be advantageous in enhancing cellular uptake of these nanoparticles and providing selective targeting. Chrysin- and selenium-loaded dendrimers are capable of induction of oxidative stress and reducing the viability of OC cells. Furthermore, they demonstrate no toxicity against normal cells that can be attributed to using folate for the functionalization of dendrimers [297].
Polymeric nanoparticles possess a core-shell structure that self-assemble in an aqueous medium. The hydrophilic shell is responsible for preserving the stability of nanoparticle, and the hydrophobic core encapsulates anti-tumor drug. Synthetic polymers, including poly (e-caprolactone) (PCL), polyglycolide (PGA), and polylactides (PLA), are applied in biomedical applications, due to their features, such as biocompatibility, high permeability, predictable degradation kinetics, etc., that are important in the field of biomedicine [298,299,300]. However, crystallinity and low biodegradation are drawbacks of PCL that can be solved using monomers. Poly (ethylene glycol) (PEG) is a safe, flexible, and hydrophilic agent approved by the Food and Drug Administration (FDA) that can be used internally in the human body [298,301,302,303]. Chrysin-loaded polymeric nanoparticles have been applied in breast cancer therapy. The results demonstrate that targeted delivery of chrysin at the tumor site by polymeric nanoparticles leads to enhanced anti-tumor activity, due to enhanced cellular uptake [304].
Nanoparticles can provide a platform for co-loading of chrysin with other natural anti-tumor compounds, such as curcumin. Briefly, curcumin is isolated from the rhizome of curcuma longa and has potent anti-tumor activity against different cancer cells [305]. Using nanoparticles can significantly enhance the bioavailability and therapeutic effects of curcumin [306]. Curcumin- and chrysin-loaded PLGA-PEG nanoparticles have been designed in CRC therapy. This co-loading exerts a synergistic effect and enhances the cytotoxicity of these phytochemicals against CRC cells [307]. Studies demonstrate that telomerase activity is associated with enhanced proliferation and invasion of cancer cells. Catalytic domain (hTERT) participates in telomerase gene overexpression that has been reported in CRC [308,309]. Chrysin- and curcumin-loaded nanoparticles effectively down-regulate the expression of hTERT in suppressing the progression of CRC cells [307]. In addition to the anti-proliferative activity via hTERT down-regulation, chrysin- and curcumin-loaded nanoparticles can suppress metastasis of cancer cells via reducing expressions of MMP-2 and MMP-9 [310].
Several homologous proteins known as tissue inhibitors of metalloproteinase (TIMPs) can regulate the activity of MMPs. TIMP-1 and TIMP-2 are capable of reducing the expression of MMP-2 and MMP-9 in suppressing metastasis and migration of cancer cells [311]. Chrysin- and curcumin-loaded nanoparticles significantly promote the expression of TIMP-1 and TIMP-2 to exert a reduction in melanoma invasion [310]. Taking everything into account, studies agree with the fact that nanoparticles can enhance the anti-tumor activity of chrysin against cancer cells [62,312,313,314,315,316]. Nanoparticles can provide a platform for the co-delivery of chrysin and other anti-tumor agents that is important in promoting its inhibitory effect against cancer cells (Figure 4) (Table 3). Further studies can focus on developing other types of nanocarriers, such as carbon nanotubes, liposomes, etc., for delivery of chrysin in cancer therapy.

8. Conclusions and Remarks

In the present review, we provided a mechanistic review of chrysin and its underlying mechanisms for anti-tumor activity [323,324,325]. Noteworthy, chrysin derivatives have also shown potential anti-tumor activity [326,327,328,329], showing that future studies can focus on chemical modification of chrysin structure in improving its bioavailability, anti-tumor activity, etc. Although chemical modification is a promising strategy in promoting the anti-tumor activity of chrysin, it seems that nanoscale delivery systems, such as polymeric nanoparticles, liposomes, solid lipid nanoparticles, etc., can also be considered in promoting cellular uptake of chrysin and enhancing its anti-tumor activity.
Chrysin affects various molecular pathways and mechanisms in cancer therapy. Apoptosis is the most well-known target of chrysin in cancer therapy, and both intrinsic and extrinsic pathways of apoptosis are induced by chrysin in cancer cells. Disrupting homeostasis of mitochondria and ER are followed by chrysin in apoptosis induction in cancer cells. Autophagy is another programmed cell death that is activated by chrysin in cancer therapy. As autophagy has a dual role in cancer, meaning it may suppress cancer progression, or may function as a pro-survival factor in promoting the proliferation of cancer cells [330,331,332,333], much attention should be directed towards the regulation of autophagy by chrysin in cancer therapy. It has been reported that chrysin induces autophagy in cancer therapy, showing the anti-tumor role of autophagy. However, more studies will reveal a relationship between chrysin and autophagy in cancer therapy. In terms of molecular pathways, oncogenic ones, such as STAT3, NF-κB, and PI3K, that are involved in cancer growth and metastasis, are suppressed upon chrysin administration. MiRs are also potential targets of chrysin in cancer therapy that their expression is regulated. Noteworthy, since studies have shown that chrysin is capable of modulating the expression of miRs, further studies can focus on evaluating the effect of chrysin on other types of non-coding RNAs, such as long non-coding RNAs (lncRNAs) and circular RNAs (circRNAs).
Another potential application of chrysin is in suppressing chemoresistance. One of the major challenges in the field of chemotherapy is the resistance of cancer cells into the inhibitory effect of currently applied chemotherapeutic agents. Chrysin induces apoptosis to sensitize cancer cells into chemotherapy. Moreover, molecular pathways, such as Nrf2, that induce chemoresistance, are suppressed via chrysin. Further studies can focus on revealing other molecular pathways, such as miRs in chemoresistance, and the role of chrysin in their regulation.
In fact, different aspects of cancer cells are affected by chrysin, including proliferation, metastasis, and chemoresistance. These inhibitory effects are mediated via affecting both molecular pathways and mechanisms that were comprehensively discussed in the main text. As poor bioavailability is one of the drawbacks of chrysin in cancer therapy, a section was allotted to examine the role of nanoparticles for promoting bioavailability and the therapeutic effects of chrysin in cancer therapy. It is worth mentioning that these results were based on in vitro and in vivo experiments. Further studies can focus on evaluating the role of chrysin in clinical studies, which is important for clinical translation of chrysin.

Author Contributions

A.P.K., A.Z. (Ali Zarrabi) and M.A. participated in the conceptualization. M.A. and A.P.K. contributed to critical editing of manuscript. Soft-ware works were performed by M.A. and A.Z. (Ali Zarrabi) Data acquisition was done by E.M. and H.A. All other authors participated in preparing first draft of manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research is funded by grants from the National Medical Research Council Singapore (grant number NMRC/OFIRG/0020/2016) to Kumar AP. The National Research Foundation Singapore and the Singapore Ministry of Education under its Research Centres of Excellence initiative to Cancer Science Institute of Singapore; National University of Singapore also supported Kumar AP for this research.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ERendoplasmic reticulum
ROSreactive oxygen species
EMTepithelial-to-mesenchymal transition
NAFLDnon-alcoholic fatty liver disease
TNF-αtumor necrosis factor-α
ILinterleukin
I/Rischemic/reperfusion
PDParkinson’s disease
TBItraumatic brain injury
Nrf2nuclear factor erythroid 2-related factor 2
EGFRepidermal growth factor receptor
VEGFvascular endothelial growth factor
HIF-1hypoxia-inducible factor-1
STAT3signal transducer and activator of transcription 3
TLRstoll-like receptors
NF-κBnuclear factor-kappaB
PCprostate cancer
UPRunfolded protein response
PERKPRKR-like ER kinase
elF2αeukaryotic translation initiation factor 2α
GRP7878 kDa glucose-regulated protein
OCovarian cancer
GCgastric cancer
5mC5-methylcytosine
5hmC5-hydroxymethylcytosine
miRmicroRNA
TGF-βtransforming growth factor-beta
HKhexokinase
HCChepatocellular carcinoma
ECMextracellular matrix
5-FU5-Fluorouracil
CRCcolorectal cancer
LC-3IIlight chain-3II
mTORmammalian target of rapamycin
PPARαPeroxisome proliferator-activated receptor alpha
CYPcytochrome C
DTXdocetaxel
FDAFood and Drug Administration
TIMPstissue inhibitors of metalloproteinases
lncRNAslong non-coding RNAs
circRNAscircular RNAs

References

  1. Siegel, R.L.; Miller, K.D.; Jemal, A. Cancer statistics, 2020. CA Cancer J. Clin. 2020, 70, 7–30. [Google Scholar] [CrossRef] [PubMed]
  2. Kuipers, E.J.; Grady, W.M.; Lieberman, D.; Seufferlein, T.; Sung, J.J.; Boelens, P.G.; van de Velde, C.J.; Watanabe, T. Colorectal cancer. Nat. Rev. Dis. Primers 2015, 1, 15065. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Das, B.; Pal, B.; Bhuyan, R.; Li, H.; Sarma, A.; Gayan, S.; Talukdar, J.; Sandhya, S.; Bhuyan, S.; Gogoi, G.; et al. MYC Regulates the HIF2α Stemness Pathway via Nanog and Sox2 to Maintain Self-Renewal in Cancer Stem Cells versus Non-Stem Cancer Cells. Cancer Res. 2019, 79, 4015–4025. [Google Scholar] [CrossRef]
  4. Fang, Z.; Li, T.; Chen, W.; Wu, D.; Qin, Y.; Liu, M.; Wu, G.; He, L.; Li, H.; Gu, H. Gab2 promotes cancer stem cell like properties and metastatic growth of ovarian cancer via downregulation of miR-200c. Exp. Cell Res. 2019, 382, 111462. [Google Scholar] [CrossRef] [PubMed]
  5. Banik, K.; Ranaware, A.M.; Harsha, C.; Nitesh, T.; Girisa, S.; Deshpande, V.; Fan, L.; Nalawade, S.P.; Sethi, G.; Kunnumakkara, A.B.; et al. Piceatannol: A natural stilbene for the prevention and treatment of cancer. Pharmacol. Res. 2020, 153, 104635. [Google Scholar] [CrossRef]
  6. Yu, Y.; Wang, Y.; Xiao, X.; Cheng, W.; Hu, L.; Yao, W.; Qian, Z.; Wu, W.; Chang, W. MiR-204 inhibits hepatocellular cancer drug resistance and metastasis through targeting NUAK1. Biochem. Cell Biol. 2019, 97, 563–570. [Google Scholar] [CrossRef]
  7. Sheng, X.; Li, Y.; Li, Y.; Liu, W.; Lu, Z.; Zhan, J.; Xu, M.; Chen, L.; Luo, X.; Cai, G.; et al. PLOD2 contributes to drug resistance in laryngeal cancer by promoting cancer stem cell-like characteristics. BMC Cancer 2019, 19, 840. [Google Scholar] [CrossRef] [Green Version]
  8. Weng, W.; Goel, A. Curcumin and colorectal cancer: An update and current perspective on this natural medicine. Semin. Cancer Biol. 2020. [Google Scholar] [CrossRef]
  9. Yang, S.-F.; Weng, C.-J.; Sethi, G.; Hu, D.-N. Natural Bioactives and Phytochemicals Serve in Cancer Treatment and Prevention. Evid.-Based Complement. Altern. Med. 2013, 2013, 698190. [Google Scholar] [CrossRef]
  10. Mishra, S.; Verma, S.S.; Rai, V.; Awasthee, N.; Chava, S.; Hui, K.M.; Kumar, A.P.; Challagundla, K.B.; Sethi, G.; Gupta, S.C. Long non-coding RNAs are emerging targets of phytochemicals for cancer and other chronic diseases. Cell. Mol. Life Sci. 2019, 76, 1947–1966. [Google Scholar] [CrossRef]
  11. Hsieh, Y.-S.; Yang, S.-F.; Sethi, G.; Hu, D.-N. Natural Bioactives in Cancer Treatment and Prevention. BioMed Res. Int. 2015, 2015, 182835. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Líšková, A.; Koklesova, L.; Samec, M.; Smejkal, K.; Samuel, S.M.; Varghese, E.; Abotaleb, M.; Biringer, K.; Kudela, E.; Danko, J.; et al. Flavonoids in Cancer Metastasis. Cancers 2020, 12, 1498. [Google Scholar] [CrossRef] [PubMed]
  13. Wu, C.-S.; Wu, S.-Y.; Chen, H.-C.; Chu, C.-A.; Tang, H.-H.; Liu, H.-S.; Hong, Y.-R.; Huang, C.-Y.; Huang, G.-C.; Su, A.C.-L. Curcumin functions as a MEK inhibitor to induce a synthetic lethal effect on KRAS mutant colorectal cancer cells receiving targeted drug regorafenib. J. Nutr. Biochem. 2019, 74, 108227. [Google Scholar] [CrossRef]
  14. Bolat, Z.B.; Islek, Z.; Demir, B.N.; Yilmaz, E.N.; Sahin, F.; Ucisik, M.H. Curcumin- and Piperine-Loaded Emulsomes as Combinational Treatment Approach Enhance the Anticancer Activity of Curcumin on HCT116 Colorectal Cancer Model. Front. Bioeng. Biotechnol. 2020, 8, 50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Zhao, R.; Du, S.; Liu, Y.; Lv, C.; Song, Y.; Chen, X.; Zhang, B.; Li, D.; Gao, S.; Cui, W.; et al. Mucoadhesive-to-penetrating controllable peptosomes-in-microspheres co-loaded with anti-miR-31 oligonucleotide and Curcumin for targeted colorectal cancer therapy. Theranostics 2020, 10, 3594–3611. [Google Scholar] [CrossRef]
  16. Deng, S.; Shanmugam, M.K.; Kumar, A.P.; Yap, C.T.; Sethi, G.; Bishayee, A. Targeting autophagy using natural compounds for cancer prevention and therapy. Cancer 2019, 125, 1228–1246. [Google Scholar] [CrossRef]
  17. Shanmugam, M.K.; Warrier, S.; Kumar, A.P.; Sethi, G.; Arfuso, F.; Shanmugam, S.W.M.K. Potential Role of Natural Compounds as Anti-Angiogenic Agents in Cancer. Curr. Vasc. Pharmacol. 2017, 15, 503–519. [Google Scholar] [CrossRef]
  18. Dai, X.; Zhang, J.; Arfuso, F.; Chinnathambi, A.; Zayed, M.; Alharbi, S.A.; Kumar, A.P.; Ahn, K.S.; Sethi, G. Targeting TNF-related apoptosis-inducing ligand (TRAIL) receptor by natural products as a potential therapeutic approach for cancer therapy. Exp. Biol. Med. 2015, 240, 760–773. [Google Scholar] [CrossRef] [Green Version]
  19. Merarchi, M.; Sethi, G.; Shanmugam, M.K.; Fan, L.; Arfuso, F.; Ahn, K.S. Role of Natural Products in Modulating Histone Deacetylases in Cancer. Molecules 2019, 24, 1047. [Google Scholar] [CrossRef] [Green Version]
  20. Kim, C.; Cho, S.K.; Kapoor, S.; Kumar, A.; Vali, S.; Abbasi, T.; Kim, S.-H.; Sethi, G.; Ahn, K.S. β-caryophyllene oxide inhibits constitutive and inducible STAT3 signaling pathway through induction of the SHP-1 protein tyrosine phosphatase. Mol. Carcinog. 2013, 53, 793–806. [Google Scholar] [CrossRef]
  21. Zhang, Q.; Zhang, Z.-Y.; Du, H.; Li, S.-Z.; Tu, R.; Jia, Y.-F.; Zheng, Z.; Song, X.-M.; Du, R.-L.; Zhang, X.-D. DUB3 deubiquitinates and stabilizes NRF2 in chemotherapy resistance of colorectal cancer. Cell Death Differ. 2019, 26, 2300–2313. [Google Scholar] [CrossRef] [PubMed]
  22. Lai, K.-C.; Chueh, F.-S.; Hsiao, Y.-T.; Cheng, Z.-Y.; Lien, J.-C.; Liu, K.-C.; Peng, S.-F.; Chung, J.-G. Gefitinib and curcumin-loaded nanoparticles enhance cell apoptosis in human oral cancer SAS cells in vitro and inhibit SAS cell xenografted tumor in vivo. Toxicol. Appl. Pharmacol. 2019, 382, 114734. [Google Scholar] [CrossRef] [PubMed]
  23. Montazerabadi, A.; Beik, J.; Irajirad, R.; Attaran, N.; Khaledi, S.; Ghaznavi, H.; Shakeri-Zadeh, A. Folate-modified and curcumin-loaded dendritic magnetite nanocarriers for the targeted thermo-chemotherapy of cancer cells. Artif. Cells Nanomed. Biotechnol. 2019, 47, 330–340. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Al-Attar, T.; Madihally, S.V. Targeted cancer treatment using a combination of siRNA-liposomes and resveratrol-electrospun fibers in co-cultures. Int. J. Pharm. 2019, 569, 118599. [Google Scholar] [CrossRef] [PubMed]
  25. Lin, C.-C.; Chin, Y.-T.; Shih, Y.-J.; Chen, Y.-R.; Chung, Y.-Y.; Lin, C.-Y.; Hsiung, C.-N.; Whang-Peng, J.; Lee, S.; Lin, H.-Y.; et al. Resveratrol antagonizes thyroid hormone-induced expression of checkpoint and proliferative genes in oral cancer cells. J. Dent. Sci. 2019, 14, 255–262. [Google Scholar] [CrossRef]
  26. Sundaram, M.K.; Raina, R.; Afroze, N.; Bajbouj, K.; Hamad, M.; Haque, S.; Hussain, A. Quercetin modulates signaling pathways and induces apoptosis in cervical cancer cells. Biosci. Rep. 2019, 39. [Google Scholar] [CrossRef] [Green Version]
  27. Prasannan, R.; Kalesh, K.A.; Shanmugam, M.K.; Nachiyappan, A.; Ramachandran, L.; Nguyen, A.H.; Kumar, A.P.; Lakshmanan, M.; Ahn, K.S.; Sethi, G. Key cell signaling pathways modulated by zerumbone: Role in the prevention and treatment of cancer. Biochem. Pharmacol. 2012, 84, 1268–1276. [Google Scholar] [CrossRef] [Green Version]
  28. Siveen, K.S.; Ahn, K.S.; Ong, T.H.; Shanmugam, M.K.; Li, F.; Yap, W.N.; Kumar, A.P.; Fong, C.W.; Tergaonkar, V.; Hui, K.M.; et al. Y-tocotrienol inhibits angiogenesis-dependent growth of human hepatocellular carcinoma through abrogation of AKT/mTOR pathway in an orthotopic mouse model. Oncotarget 2014, 5, 1897–1911. [Google Scholar] [CrossRef] [Green Version]
  29. Siveen, K.S.; Mustafa, N.; Li, F.; Kannaiyan, R.; Ahn, K.S.; Kumar, A.P.; Chng, W.-J.; Sethi, G. Thymoquinone overcomes chemoresistance and enhances the anticancer effects of bortezomib through abrogation of NF-κB regulated gene products in multiple myeloma xenograft mouse model. Oncotarget 2013, 5, 634–648. [Google Scholar] [CrossRef] [Green Version]
  30. Wu, M.; Zhang, P. EGFR-mediated autophagy in tumourigenesis and therapeutic resistance. Cancer Lett. 2019, 469, 207–216. [Google Scholar] [CrossRef]
  31. Kim, S.-H.; Yoo, E.-S.; Woo, J.-S.; Han, S.-H.; Lee, J.-H.; Jung, S.-H.; Kim, H.-J.; Jung, J.-Y. Antitumor and apoptotic effects of quercetin on human melanoma cells involving JNK/P38 MAPK signaling activation. Eur. J. Pharmacol. 2019, 860, 172568. [Google Scholar] [CrossRef] [PubMed]
  32. Tewari, D.; Nabavi, S.F.; Nabavi, S.M.; Sureda, A.; Farooqi, A.A.; Atanasov, A.G.; Vacca, R.A.; Sethi, G.; Fimognari, C. Targeting activator protein 1 signaling pathway by bioactive natural agents: Possible therapeutic strategy for cancer prevention and intervention. Pharmacol. Res. 2017, 128, 366–375. [Google Scholar] [CrossRef] [PubMed]
  33. Lee, J.H.; Chiang, S.Y.; Nam, D.; Chung, W.-S.; Lee, J.; Na, Y.-S.; Sethi, G.; Ahn, K.S. Capillarisin inhibits constitutive and inducible STAT3 activation through induction of SHP-1 and SHP-2 tyrosine phosphatases. Cancer Lett. 2014, 345, 140–148. [Google Scholar] [CrossRef] [PubMed]
  34. Patel, S.M.; Venkata, K.C.N.; Bhattacharyya, P.; Sethi, G.; Bishayee, A. Potential of neem (Azadirachta indica L.) for prevention and treatment of oncologic diseases. Semin. Cancer Biol. 2016, 40–41, 100–115. [Google Scholar] [CrossRef]
  35. Ko, J.; Yang, M.H.; Baek, S.H.; Nam, D.; Jung, S.H.; Ahn, K.S. Theacrine attenuates epithelial mesenchymal transition in human breast cancer MDA-MB-231 cells. Phytother. Res. 2019, 33, 1934–1942. [Google Scholar] [CrossRef]
  36. Cheng, J.-T.; Wang, L.; Wang, H.; Tang, F.R.; Cai, W.; Sethi, G.; Xin, H.-W.; Ma, Z. Insights into Biological Role of LncRNAs in Epithelial-Mesenchymal Transition. Cells 2019, 8, 1178. [Google Scholar] [CrossRef] [Green Version]
  37. Loh, C.-Y.; Chai, J.Y.; Tang, T.F.; Wong, W.; Sethi, G.; Shanmugam, M.K.; Chong, P.P.; Looi, C.Y. The E-Cadherin and N-Cadherin Switch in Epithelial-to-Mesenchymal Transition: Signaling, Therapeutic Implications, and Challenges. Cells 2019, 8, 1118. [Google Scholar] [CrossRef] [Green Version]
  38. Yuan, L.; Zhou, M.; Huang, D.; Wasan, H.S.; Zhang, K.; Sun, L.; Huang, H.; Ma, S.; Shen, M.; Ruan, S. Resveratrol inhibits the invasion and metastasis of colon cancer through reversal of epithelial- mesenchymal transition via the AKT/GSK-3β/Snail signaling pathway. Mol. Med. Rep. 2019, 20, 2783–2795. [Google Scholar] [CrossRef]
  39. Lee, J.H.; Mohan, C.D.; Deivasigamani, A.; Jung, Y.Y.; Rangappa, S.; Basappa, S.; Chinnathambi, A.; Alahmadi, T.A.; Alharbi, S.A.; Garg, M.; et al. Brusatol suppresses STAT3-driven metastasis by downregulating epithelial-mesenchymal transition in hepatocellular carcinoma. J. Adv. Res. 2020. [Google Scholar] [CrossRef]
  40. Lee, J.H.; Mohan, C.D.; Shanmugam, M.K.; Rangappa, S.; Sethi, G.; Siveen, K.S.; Chinnathambi, A.; Alahmadi, T.A.; Alharbi, S.A.; Basappa, S.; et al. Vitexin abrogates invasion and survival of hepatocellular carcinoma cells through targeting STAT3 signaling pathway. Biochimie 2020, 175, 58–68. [Google Scholar] [CrossRef]
  41. Aggarwal, V.; Tuli, H.S.; Kaur, J.; Aggarwal, D.; Parashar, G.; Parashar, N.C.; Kulkarni, S.; Kaur, G.; Sak, K.; Kumar, M.; et al. Garcinol Exhibits Anti-Neoplastic Effects by Targeting Diverse Oncogenic Factors in Tumor Cells. Biomedicines 2020, 8, 103. [Google Scholar] [CrossRef] [PubMed]
  42. Shahzadi, I.; Ali, Z.; Baek, S.H.; Mirza, B.; Ahn, K.S. Assessment of the Antitumor Potential of Umbelliprenin, a Naturally Occurring Sesquiterpene Coumarin. Biomedicines 2020, 8, 126. [Google Scholar] [CrossRef] [PubMed]
  43. Henamayee, S.; Banik, K.; Sailo, B.L.; Shabnam, B.; Harsha, C.; Srilakshmi, S.; Naidu, V.; Baek, S.H.; Ahn, K.S.; Kunnumakkara, A.B. Therapeutic Emergence of Rhein as a Potential Anticancer Drug: A Review of Its Molecular Targets and Anticancer Properties. Molecules 2020, 25, 2278. [Google Scholar] [CrossRef]
  44. Hwang, S.T.; Um, J.-Y.; Chinnathambi, A.; Alharbi, S.A.; Narula, A.S.; Namjoshi, O.A.; Blough, B.E.; Ahn, K.S. Evodiamine Mitigates Cellular Growth and Promotes Apoptosis by Targeting the c-Met Pathway in Prostate Cancer Cells. Molecules 2020, 25, 1320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Naz, I.; Ramchandani, S.; Khan, R.A.; Yang, M.H.; Ahn, K.S. Anticancer Potential of Raddeanin A, a Natural Triterpenoid Isolated from Anemone raddeana Regel. Molecules 2020, 25, 1035. [Google Scholar] [CrossRef] [Green Version]
  46. Pingili, R.; Pawar, A.K.; Challa, S.R.; Kodali, T.; Koppula, S.; Toleti, V. A comprehensive review on hepatoprotective and nephroprotective activities of chrysin against various drugs and toxic agents. Chem. Interact. 2019, 308, 51–60. [Google Scholar] [CrossRef]
  47. Angelopoulou, E.; Pyrgelis, E.-S.; Piperi, C. Neuroprotective potential of chrysin in Parkinson’s disease: Molecular mechanisms and clinical implications. Neurochem. Int. 2020, 132, 104612. [Google Scholar] [CrossRef]
  48. Naz, S.; Imran, M.; Rauf, A.; Orhan, I.E.; Shariati, M.A.; Shahbaz, M.; Qaisrani, T.B.; Shah, Z.A.; Plygun, S.; Heydari, M.; et al. Chrysin: Pharmacological and therapeutic properties. Life Sci. 2019, 235, 116797. [Google Scholar] [CrossRef]
  49. Nabavi, S.F.; Braidy, N.; Habtemariam, S.; Orhan, I.E.; Daglia, M.; Manayi, A.; Gortzi, O.; Nabavi, S.M. Neuroprotective effects of chrysin: From chemistry to medicine. Neurochem. Int. 2015, 90, 224–231. [Google Scholar] [CrossRef]
  50. Pietta, P.-G. Flavonoids as Antioxidants. J. Nat. Prod. 2000, 63, 1035–1042. [Google Scholar] [CrossRef]
  51. Xu, H.; Luo, J.; Huang, J.; Wen, Q. Flavonoids intake and risk of type 2 diabetes mellitus. Medicine 2018, 97, e0686. [Google Scholar] [CrossRef] [PubMed]
  52. Bajgai, S.P.; Prachyawarakorn, V.; Mahidol, C.; Ruchirawat, S.; Kittakoop, P. Hybrid flavan-chalcones, aromatase and lipoxygenase inhibitors, from Desmos cochinchinensis. Phytochemistry 2011, 72, 2062–2067. [Google Scholar] [CrossRef] [PubMed]
  53. Escuredo, O.; Silva, L.R.; Valentão, P.; Seijo, M.C.; Andrade, P.B. Assessing Rubus honey value: Pollen and phenolic compounds content and antibacterial capacity. Food Chem. 2012, 130, 671–678. [Google Scholar] [CrossRef]
  54. Hadjmohammadi, M.R.; Nazari, S.S.S.J. Separation optimization of quercetin, hesperetin and chrysin in honey by micellar liquid chromatography and experimental design. J. Sep. Sci. 2010, 33, 3144–3151. [Google Scholar] [CrossRef]
  55. Canini, A.; Pichichero, E.; Cicconi, R.; Mattei, M.; Muzi, M.G. Acacia honey and chrysin reduce proliferation of melanoma cells through alterations in cell cycle progression. Int. J. Oncol. 2010, 37, 973–981. [Google Scholar] [CrossRef] [Green Version]
  56. Kalogeropoulos, N.; Yanni, A.E.; Koutrotsios, G.; Aloupi, M. Bioactive microconstituents and antioxidant properties of wild edible mushrooms from the island of Lesvos, Greece. Food Chem. Toxicol. 2013, 55, 378–385. [Google Scholar] [CrossRef]
  57. Balta, C.; Herman, H.; Boldura, O.; Gasca, I.; Rosu, M.; Ardelean, A.; Hermenean, A. Chrysin attenuates liver fibrosis and hepatic stellate cell activation through TGF-β/Smad signaling pathway. Chem. Interact. 2015, 240, 94–101. [Google Scholar] [CrossRef]
  58. Mani, R.; Natesan, V. Chrysin: Sources, beneficial pharmacological activities, and molecular mechanism of action. Phytochemistry 2018, 145, 187–196. [Google Scholar] [CrossRef]
  59. Galijatovic, A.; Otake, Y.; Walle, U.K.; Walle, T. Extensive metabolism of the flavonoid chrysin by human Caco-2 and Hep G2 cells. Xenobiotica 1999, 29, 1241–1256. [Google Scholar] [CrossRef]
  60. Walle, T.; Otake, Y.; Brubaker, J.A.; Walle, U.K.; Halushka, P.V. Disposition and metabolism of the flavonoid chrysin in normal volunteers. Br. J. Clin. Pharmacol. 2001, 51, 143–146. [Google Scholar]
  61. Walle, U.; Galijatovic, A.; Walle, T. Transport of the flavonoid chrysin and its conjugated metabolites by the human intestinal cell line Caco-2. Biochem. Pharmacol. 1999, 58, 431–438. [Google Scholar] [CrossRef]
  62. Ferrado, J.B.; Perez, A.A.; Visentini, F.F.; Islan, G.A.; Castro, G.R.; Santiago, L.G. Formation and characterization of self-assembled bovine serum albumin nanoparticles as chrysin delivery systems. Colloids Surf. B Biointerfaces 2019, 173, 43–51. [Google Scholar] [CrossRef] [PubMed]
  63. Belhan, S.; Yıldırım, S.; Karasu, A.; Kömüroğlu, A.U.; Özdek, U. Investigation of the protective role of chrysin within the framework of oxidative and inflammatory markers in experimental testicular ischaemia/reperfusion injury in rats. Andrologia 2020, 13714. [Google Scholar] [CrossRef] [PubMed]
  64. Sassi, A.; Boubaker, J.; Loussaief, A.; Jomaa, K.; Ghedira, K.; Chekir-Ghedira, L. Protective Effect of Chrysin, a Dietary Flavone against Genotoxic and Oxidative Damage Induced by Mitomycin C in Balb/C Mice. Nutr. Cancer 2020, 1–10. [Google Scholar] [CrossRef] [PubMed]
  65. Yao, J.; Jiang, M.; Zhang, Y.; Liu, X.; Li, Y.; Feng, G. Chrysin alleviates allergic inflammation and airway remodeling in a murine model of chronic asthma. Int. Immunopharmacol. 2016, 32, 24–31. [Google Scholar] [CrossRef]
  66. Lee, E.-J.; Kang, M.-K.; Kim, Y.-H.; Kim, D.Y.; Oh, H.; Kim, S.-I.; Oh, S.Y.; Kang, Y.-H. Dietary Chrysin Suppresses Formation of Actin Cytoskeleton and Focal Adhesion in AGE-Exposed Mesangial Cells and Diabetic Kidney: Role of Autophagy. Nutrients 2019, 11, 127. [Google Scholar] [CrossRef] [Green Version]
  67. Bortolotto, V.C.; Araujo, S.M.; Pinheiro, F.C.; Poetini, M.R.; De Paula, M.T.; Meichtry, L.B.; De Almeida, F.P.; Musachio, E.A.S.; Guerra, G.P.; Prigol, M. Modulation of glutamate levels and Na+, K+-ATPase activity contributes to the chrysin memory recovery in hypothyroidism mice. Physiol. Behav. 2020, 222, 112892. [Google Scholar] [CrossRef]
  68. Song, Y.; Wu, W.; Sheng, L.; Jiang, B.; Li, X.; Cai, K. Chrysin ameliorates hepatic steatosis induced by a diet deficient in methionine and choline by inducing the secretion of hepatocyte nuclear factor 4α-dependent very low-density lipoprotein. J. Biochem. Mol. Toxicol. 2020, 34, e22497. [Google Scholar] [CrossRef]
  69. Yang, M.; Xiong, J.; Zou, Q.; Wang, D.-D.; Huang, C. Chrysin attenuates interstitial fibrosis and improves cardiac function in a rat model of acute myocardial infarction. J. Mol. Histol. 2018, 49, 555–565. [Google Scholar] [CrossRef]
  70. Choi, J.H.; Yun, J.W. Chrysin induces brown fat–like phenotype and enhances lipid metabolism in 3T3-L1 adipocytes. Nutrients 2016, 32, 1002–1010. [Google Scholar] [CrossRef]
  71. Pai, S.A.; Munshi, R.P.; Panchal, F.H.; Gaur, I.-S.; Juvekar, A.R. Chrysin ameliorates nonalcoholic fatty liver disease in rats. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2019, 392, 1617–1628. [Google Scholar] [CrossRef] [PubMed]
  72. Mohammadi, A.; Kazemi, S.; Hosseini, M.; Varzi, H.N.; Feyzi, F.; Morakabati, P.; Moghadamnia, A.A.; Mohammadi, A. Chrysin Effect in Prevention of Acetaminophen-Induced Hepatotoxicity in Rat. Chem. Res. Toxicol. 2019, 32, 2329–2337. [Google Scholar] [CrossRef] [PubMed]
  73. Xu, M.; Shi, H.; Liu, D. Chrysin protects against renal ischemia reperfusion induced tubular cell apoptosis and inflammation in mice. Exp. Ther. Med. 2019, 17, 2256–2262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Melekoglu, R.; Çiftçi, O.; Eraslan, S.; Alan, S.; Başak, N. The Protective Effects of Glycyrrhetinic Acid and Chrysin against Ischemia-Reperfusion Injury in Rat Ovaries. BioMed Res. Int. 2018, 2018, 5421308. [Google Scholar] [CrossRef]
  75. Yao, Y.; Chen, L.; Xiao, J.; Wang, C.; Jiang, W.; Zhang, R.; Hao, J. Chrysin Protects against Focal Cerebral Ischemia/Reperfusion Injury in Mice through Attenuation of Oxidative Stress and Inflammation. Int. J. Mol. Sci. 2014, 15, 20913–20926. [Google Scholar] [CrossRef] [Green Version]
  76. Saliba-Gustafsson, P.; Pedrelli, M.; Gertow, K.; Werngren, O.; Janas, V.; Pourteymour, S.; Baldassarre, D.; Tremoli, E.; Veglia, F.; Rauramaa, E.; et al. Subclinical atherosclerosis and its progression are modulated by PLIN2 through a feed-forward loop between LXR and autophagy. J. Intern. Med. 2019, 286, 660–675. [Google Scholar] [CrossRef] [Green Version]
  77. Krishnamoorthy, A.; Sevanan, M.; Mani, S.; Balu, M.; Balaji, S.; Ramajayan, P. Chrysin restores MPTP induced neuroinflammation, oxidative stress and neurotrophic factors in an acute Parkinson’s disease mouse model. Neurosci. Lett. 2019, 709, 134382. [Google Scholar] [CrossRef]
  78. Rashno, M.; Ghaderi, S.; Nesari, A.; Khorsandi, L.; Farbood, Y.; Sarkaki, A. Chrysin attenuates traumatic brain injury-induced recognition memory decline, and anxiety/depression-like behaviors in rats: Insights into underlying mechanisms. Psychopharmacology 2020, 237, 1–13. [Google Scholar] [CrossRef]
  79. Dong, F.; Zhang, J.; Zhu, S.; Lan, T.; Yang, J.; Li, L. Chrysin Alleviates Chronic Hypoxia–Induced Pulmonary Hypertension by Reducing Intracellular Calcium Concentration in Pulmonary Arterial Smooth Muscle Cells. J. Cardiovasc. Pharmacol. 2019, 74, 426–435. [Google Scholar] [CrossRef]
  80. Li, H.-J.; Wu, N.-L.; Pu, C.-M.; Hsiao, C.-Y.; Chang, D.-C.; Hung, C.-F. Chrysin alleviates imiquimod-induced psoriasis-like skin inflammation and reduces the release of CCL20 and antimicrobial peptides. Sci. Rep. 2020, 10, 2932. [Google Scholar] [CrossRef] [Green Version]
  81. Koc, F.; Tekeli, M.Y.; Kanbur, M.; Karayigit, M.Ö.; Liman, B.C. The effects of chrysin on lipopolysaccharide-induced sepsis in rats. J. Food Biochem. 2020, e13359. [Google Scholar] [CrossRef] [PubMed]
  82. Del Fabbro, L.; De Gomes, M.G.; Souza, L.C.; Goes, A.R.; Boeira, S.P.; Oliveira, M.S.; Furian, A.F.; Jesse, C.R. Chrysin suppress immune responses and protects from experimental autoimmune encephalomyelitis in mice. J. Neuroimmunol. 2019, 335, 577007. [Google Scholar] [CrossRef] [PubMed]
  83. Wei, Y.; Zheng, Q.; Tang, G.; Song, C.; Wang, G.; Zhang, Y.; Xiao, Y.; Zeng, X.; Wang, Z.; Xiao, J.; et al. Synthesis and Anti-Thyroid Cancer Effect of Iodo-Chrysin Derivatives. Med. Chem. 2015, 12, 441–447. [Google Scholar] [CrossRef] [PubMed]
  84. Sulaiman, G.M.; Jabir, M.S.; Hameed, A.H. Nanoscale modification of chrysin for improved of therapeutic efficiency and cytotoxicity. Artif. Cells Nanomed. Biotechnol. 2018, 46, 708–720. [Google Scholar] [CrossRef] [PubMed]
  85. Andrade, N.; Andrade, S.; Silva, C.; Rodrigues, I.; Guardão, L.; Guimarães, J.T.; Keating, E.; Martel, F. Chronic consumption of the dietary polyphenol chrysin attenuates metabolic disease in fructose-fed rats. Eur. J. Nutr. 2019, 59, 151–165. [Google Scholar] [CrossRef] [PubMed]
  86. Prajit, R.; Sritawan, N.; Suwannakot, K.; Naewla, S.; Aranarochana, A.; Sirichoat, A.; Pannangrong, W.; Wigmore, P.; Welbat, J.U. Chrysin Protects against Memory and Hippocampal Neurogenesis Depletion in D-Galactose-Induced Aging in Rats. Nutrients 2020, 12, 1100. [Google Scholar] [CrossRef]
  87. Shooshtari, M.K.; Sarkaki, A.; Mansouri, S.M.T.; Badavi, M.; Khorsandi, L.; Dehcheshmeh, M.G.; Farbood, Y. Protective effects of Chrysin against memory impairment, cerebral hyperemia and oxidative stress after cerebral hypoperfusion and reperfusion in rats. Metab. Brain Dis. 2019, 35, 401–412. [Google Scholar] [CrossRef]
  88. Khezri, S.; Sabzalipour, T.; Jahedsani, A.; Azizian, S.; Atashbar, S.; Salimi, A. Chrysin ameliorates aluminum p hosphide-induced oxidative stress and mitochondrial damages in rat cardiomyocytes and isolated mitochondria. Environ. Toxicol. 2020. [Google Scholar] [CrossRef]
  89. Baykalir, B.G.; Arslan, A.S.; Mutlu, S.I.; Ak, T.P.; Seven, I.; Seven, P.T.; Yaman, M.; Gul, H.F. The protective effect of chrysin against carbon tetrachloride-induced kidney and liver tissue damage in rats. Int. J. Vitam. Nutr. Res. 2020, 1–12. [Google Scholar] [CrossRef]
  90. Temel, Y.; Kucukler, S.; Yıldırım, S.; Caglayan, C.; Kandemir, F.M. Protective effect of chrysin on cyclophosphamide-induced hepatotoxicity and nephrotoxicity via the inhibition of oxidative stress, inflammation, and apoptosis. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2019, 393, 325–337. [Google Scholar] [CrossRef]
  91. Liao, Z.-Y.; Liang, I.-C.; Li, H.-J.; Wu, C.-C.; Lo, H.-M.; Chang, D.-C.; Hung, C.-F. Chrysin Inhibits High Glucose-Induced Migration on Chorioretinal Endothelial Cells via VEGF and VEGFR Down-Regulation. Int. J. Mol. Sci. 2020, 21, 5541. [Google Scholar] [CrossRef] [PubMed]
  92. Wojnar, W.; Zych, M.; Borymski, S.; Kaczmarczyk-Sedlak, I. Chrysin Reduces Oxidative Stress but Does Not Affect Polyol Pathway in the Lenses of Type 1 Diabetic Rats. Antioxidants 2020, 9, 160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Fagundes, F.L.; Piffer, G.D.M.; Périco, L.L.; Rodrigues, V.P.; Lima, C.A.H.; Santos, R.D.C.D. Chrysin Modulates Genes Related to Inflammation, Tissue Remodeling, and Cell Proliferation in the Gastric Ulcer Healing. Int. J. Mol. Sci. 2020, 21, 760. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. DeSantis, C.E.; Bryan, L.; Jemal, A. Breast cancer statistics, 2019. CA Cancer J. Clin. 2019, 69, 438–451. [Google Scholar] [CrossRef] [PubMed]
  95. Wang, C.; Kar, S.; Lai, X.; Cai, W.; Arfuso, F.; Sethi, G.; Lobie, P.E.; Goh, B.C.; Lim, L.H.K.; Hartman, M.; et al. Triple negative breast cancer in Asia: An insider’s view. Cancer Treat. Rev. 2018, 62, 29–38. [Google Scholar] [CrossRef] [PubMed]
  96. Jia, L.Y.; Shanmugam, M.K.; Sethi, G.; Bishayee, A. Potential role of targeted therapies in the treatment of triple-negative breast cancer. Anti-Cancer Drugs 2016, 27, 147–155. [Google Scholar] [CrossRef]
  97. Haque, M.; Desai, K.V. Pathways to Endocrine Therapy Resistance in Breast Cancer. Front. Endocrinol. 2019, 10, 573. [Google Scholar] [CrossRef]
  98. Pandey, K.; An, H.-J.; Kim, S.K.; Lee, S.A.; Kim, S.; Lim, S.M.; Kim, G.M.; Sohn, J.; Moon, Y.W. Molecular mechanisms of resistance to CDK4/6 inhibitors in breast cancer: A review. Int. J. Cancer 2019, 145, 1179–1188. [Google Scholar] [CrossRef] [Green Version]
  99. Bhuvanalakshmi, G.; Rangappa, K.S.; Dharmarajan, A.; Sethi, G.; Kumar, A.P.; Warrier, S. Basappa Breast Cancer Stem-Like Cells Are Inhibited by Diosgenin, a Steroidal Saponin, by the Attenuation of the Wnt β-Catenin Signaling via the Wnt Antagonist Secreted Frizzled Related Protein-4. Front. Pharmacol. 2017, 8, 124. [Google Scholar] [CrossRef] [Green Version]
  100. Chen, L.; Yuan, Y.; Kar, S.; Kanchi, M.M.; Arora, S.; Kim, J.E.; Koh, P.F.; Yousef, E.; Samy, R.P.; Shanmugam, M.K.; et al. PPARγ Ligand–induced Annexin A1 Expression Determines Chemotherapy Response via Deubiquitination of Death Domain Kinase RIP in Triple-negative Breast Cancers. Mol. Cancer Ther. 2017, 16, 2528–2542. [Google Scholar] [CrossRef] [Green Version]
  101. Wang, L.; Wang, C.; Tao, Z.; Zhao, L.; Zhu, Z.; Wu, W.; He, Y.; Chen, H.; Zheng, B.; Huang, X.; et al. Curcumin derivative WZ35 inhibits tumor cell growth via ROS-YAP-JNK signaling pathway in breast cancer. J. Exp. Clin. Cancer Res. 2019, 38, 460. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Zhang, W.; Jiang, H.; Chen, Y.; Ren, F. Resveratrol chemosensitizes adriamycin-resistant breast cancer cells by modulating miR-122-5p. J. Cell. Biochem. 2019, 120, 16283–16292. [Google Scholar] [CrossRef] [PubMed]
  103. Shanmugam, M.K.; Ahn, K.S.; Hsu, A.; Woo, C.C.; Yuan, Y.; Tan, K.H.B.; Chinnathambi, A.; Alahmadi, T.A.; Alharbi, S.A.; Koh, A.P.F.; et al. Thymoquinone Inhibits Bone Metastasis of Breast Cancer Cells Through Abrogation of the CXCR4 Signaling Axis. Front. Pharmacol. 2018, 9, 1294. [Google Scholar] [CrossRef] [PubMed]
  104. Liu, L.; Ahn, K.S.; Shanmugam, M.K.; Wang, H.; Shen, H.; Arfuso, F.; Chinnathambi, A.; Alharbi, S.A.; Chang, Y.; Sethi, G.; et al. Oleuropein induces apoptosis via abrogating NF-κB activation cascade in estrogen receptor–negative breast cancer cells. J. Cell. Biochem. 2018, 120, 4504–4513. [Google Scholar] [CrossRef]
  105. Maasomi, Z.J.; Soltanahmadid, Y.P.; Dadashpour, M.; Alipour, S.; Abolhasani, S.; Zarghami, N. Synergistic Anticancer Effects of Silibinin and Chrysin in T47D Breast Cancer Cells. Asian Pac. J. Cancer Prev. 2017, 18, 1283–1287. [Google Scholar]
  106. Kolibaba, K.S.; Druker, B.J. Protein tyrosine kinases and cancer. Biochim. Biophys. Acta (BBA) Bioenergy 1997, 1333, F217–F248. [Google Scholar] [CrossRef]
  107. Lacouture, M.E. Mechanisms of cutaneous toxicities to EGFR inhibitors. Nat. Rev. Cancer 2006, 6, 803–812. [Google Scholar] [CrossRef]
  108. Choowongkomon, K.; Sawatdichaikul, O.; Songtawee, N.; Limtrakul, J. Receptor-Based Virtual Screening of EGFR Kinase Inhibitors from the NCI Diversity Database. Molecules 2010, 15, 4041–4054. [Google Scholar] [CrossRef] [Green Version]
  109. Debnath, S.; Kanakaraju, M.; Islam, M.; Yeeravalli, R.; Sen, D.; Das, A. In silico design, synthesis and activity of potential drug-like chrysin scaffold-derived selective EGFR inhibitors as anticancer agents. Comput. Biol. Chem. 2019, 83, 107156. [Google Scholar] [CrossRef]
  110. Yun, C.W.; Lee, J.H.; Lee, S.H. Hypoxia-induced PGC-1α Regulates Mitochondrial Function and Tumorigenesis of Colorectal Cancer Cells. Anticancer Res. 2019, 39, 4865–4876. [Google Scholar] [CrossRef]
  111. Ren, S.; Liu, J.; Feng, Y.; Li, Z.; He, L.; Li, L.; Cao, X.; Wang, Z.; Zhang, Y. Knockdown of circDENND4C inhibits glycolysis, migration and invasion by up-regulating miR-200b/c in breast cancer under hypoxia. J. Exp. Clin. Cancer Res. 2019, 38, 388. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Multhoff, G.; Vaupel, P. Hypoxia Compromises Anti-Cancer Immune Responses. Adv. Exp. Med. Biol. 2020, 131–143. [Google Scholar] [CrossRef]
  113. Brizel, D.M.; Scully, S.P.; Harrelson, J.M.; Layfield, L.J.; Bean, J.M.; Prosnitz, L.R.; Dewhirst, M.W. Tumor oxygenation predicts for the likelihood of distant metastases in human soft tissue sarcoma. Cancer Res. 1996, 56, 941–943. [Google Scholar] [PubMed]
  114. Hockel, M.; Schlenger, K.; Aral, B.; Mitze, M.; Schaffer, U.; Vaupel, P. Association between tumor hypoxia and malignant progression in advanced cancer of the uterine cervix. Cancer Res. 1996, 56, 4509–4515. [Google Scholar]
  115. Minet, E.; Michel, G.; Mottet, D.; Raes, M.; Michiels, C. Transduction pathways involved in Hypoxia-Inducible Factor-1 phosphorylation and activation. Free Radic. Biol. Med. 2001, 31, 847–855. [Google Scholar] [CrossRef]
  116. Jung, J.E.; Lee, H.-G.; Cho, I.-H.; Chung, D.H.; Yoon, S.-H.; Yang, Y.M.; Lee, J.W.; Choi, S.; Park, J.-W.; Ye, S.-K.; et al. STAT3 is a potential modulator of HIF-1-mediated VEGF expression in human renal carcinoma cells. FASEB J. 2005, 19, 1296–1298. [Google Scholar] [CrossRef]
  117. Barr, M.P.; Bouchier-Hayes, D.J.; Harmey, J.J. Vascular endothelial growth factor is an autocrine survival factor for breast tumour cells under hypoxia. Int. J. Oncol. 2008, 32, 41–48. [Google Scholar] [CrossRef] [Green Version]
  118. Zhang, J.; Lu, A.; Beech, D.; Jiang, B.; Lu, Y. Suppression of breast cancer metastasis through the inhibition of VEGF-mediated tumor angiogenesis. Cancer Ther. 2007, 5, 273–286. [Google Scholar]
  119. Lirdprapamongkol, K.; Sakurai, H.; Abdelhamed, S.; Yokoyama, S.; Maruyama, T.; Athikomkulchai, S.; Viriyaroj, A.; Awale, S.; Yagita, H.; Ruchirawat, S.; et al. A flavonoid chrysin suppresses hypoxic survival and metastatic growth of mouse breast cancer cells. Oncol. Rep. 2013, 30, 2357–2364. [Google Scholar] [CrossRef] [Green Version]
  120. Park, J.; Kim, Y.; Park, E.H.; Lee, S.; Kim, H.; Kim, A.; Lee, S.B.; Shim, S.; Jang, H.; Myung, J.K.; et al. Effects of metformin and phenformin on apoptosis and epithelial-mesenchymal transition in chemoresistant rectal cancer. Cancer Sci. 2019, 110, 2834–2845. [Google Scholar] [CrossRef]
  121. Xue, J.; Li, L.; Li, N.; Li, F.; Qin, X.; Li, T.; Liu, M. Metformin suppresses cancer cell growth in endometrial carcinoma by inhibiting PD-L1. Eur. J. Pharmacol. 2019, 859, 172541. [Google Scholar] [CrossRef] [PubMed]
  122. Rasouli, S.; Zarghami, N. Synergistic Growth Inhibitory Effects of Chrysin and Metformin Combination on Breast Cancer Cells through hTERT and Cyclin D1 Suppression. Asian Pac. J. Cancer Prev. 2018, 19, 977–982. [Google Scholar] [PubMed]
  123. International Agency for Research on Cancer. Chromium, Nickel and Welding. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans; IARC: Lyon, France, 1990; Volume 49. [Google Scholar]
  124. Jung, Y.Y.; Shanmugam, M.K.; Narula, A.S.; Kim, C.; Lee, J.H.; Namjoshi, O.A.; Blough, B.E.; Sethi, G.; Ahn, K.S. Oxymatrine Attenuates Tumor Growth and Deactivates STAT5 Signaling in a Lung Cancer Xenograft Model. Cancers 2019, 11, 49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Lee, J.H.; Kim, C.; Lee, S.-G.; Yang, W.M.; Um, J.-Y.; Sethi, G.; Ahn, K.S. Ophiopogonin D modulates multiple oncogenic signaling pathways, leading to suppression of proliferation and chemosensitization of human lung cancer cells. Phytomedicine 2018, 40, 165–175. [Google Scholar] [CrossRef]
  126. Kuo, C.-Y.; Wong, R.-H.; Lin, J.-Y.; Lai, J.-C.; Lee, H. Accumulation of Chromium and Nickel Metals in Lung Tumors from Lung Cancer Patients in Taiwan. J. Toxicol. Environ. Health Part A 2006, 69, 1337–1344. [Google Scholar] [CrossRef]
  127. Huang, H.-H.; Huang, J.-Y.; Lung, C.-C.; Wu, C.-L.; Ho, C.-C.; Sun, Y.-H.; Ko, P.-C.; Su, S.-Y.; Chen, S.-C.; Wang, B.-Y. Cell-type specificity of lung cancer associated with low-dose soil heavy metal contamination in Taiwan: An ecological study. BMC Public Health 2013, 13, 330. [Google Scholar] [CrossRef] [Green Version]
  128. Wu, C.-H.; Tang, S.C.; Wang, P.-H.; Lee, H.; Ko, J.-L. Nickel-induced Epithelial-Mesenchymal Transition by Reactive Oxygen Species Generation and E-cadherin Promoter Hypermethylation. J. Biol. Chem. 2012, 287, 25292–25302. [Google Scholar] [CrossRef] [Green Version]
  129. Chiou, Y.-H.; Liou, S.-H.; Wong, R.-H.; Chen, C.-Y.; Lee, H. Nickel may contribute to EGFR mutation and synergistically promotes tumor invasion in EGFR-mutated lung cancer via nickel-induced microRNA-21 expression. Toxicol. Lett. 2015, 237, 46–54. [Google Scholar] [CrossRef]
  130. Guo, H.; Deng, H.; Cui, H.; Peng, X.; Fang, J.; Zuo, Z.; Deng, J.; Wang, X.; Wu, B.; Chen, K. Nickel chloride (NiCl2)-caused inflammatory responses via activation of NF-κB pathway and reduction of anti-inflammatory mediator expression in the kidney. Oncotarget 2015, 6, 28607–28620. [Google Scholar] [CrossRef] [Green Version]
  131. Basith, S.; Manavalan, B.; Yoo, T.H.; Kim, S.G.; Choi, S. Roles of toll-like receptors in Cancer: A double-edged sword for defense and offense. Arch. Pharmacal Res. 2012, 35, 1297–1316. [Google Scholar] [CrossRef]
  132. Yesudhas, D.; Gosu, V.; Anwar, M.A.; Choi, S. Multiple Roles of Toll-Like Receptor 4 in Colorectal Cancer. Front. Immunol. 2014, 5, 334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Puar, Y.R.; Shanmugam, M.K.; Fan, L.; Arfuso, F.; Sethi, G.; Tergaonkar, V. Evidence for the Involvement of the Master Transcription Factor NF-κB in Cancer Initiation and Progression. Biomedicines 2018, 6, 82. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Wu, T.C.; Chan, S.T.; Chang, C.N.; Yu, P.S.; Chuang, C.H.; Yeh, S.L. Quercetin and chrysin inhibit nickel-induced invasion and migration by downregulation of TLR4/NF-κB signaling in A549 cells. Chem. Biol. Interact. 2018, 292, 101–109. [Google Scholar] [CrossRef] [PubMed]
  135. Siegel, R.L.; Miller, K.D.; Jemal, A. Cancer statistics, 2016. CA Cancer J. Clin. 2016, 66, 7–30. [Google Scholar] [CrossRef] [Green Version]
  136. Lee, J.H.; Kim, C.; Baek, S.H.; Ko, J.-H.; Lee, S.G.; Yang, W.M.; Um, J.-Y.; Sethi, G.; Ahn, K.S. Capsazepine inhibits JAK/STAT3 signaling, tumor growth, and cell survival in prostate cancer. Oncotarget 2016, 8, 17700–17711. [Google Scholar] [CrossRef]
  137. Zhang, J.; Ahn, K.S.; Kim, C.; Shanmugam, M.K.; Siveen, K.S.; Arfuso, F.; Samym, R.P.; Deivasigamanim, A.; Lim, L.H.K.; Wang, L.; et al. Nimbolide-Induced Oxidative Stress Abrogates STAT3 Signaling Cascade and Inhibits Tumor Growth in Transgenic Adenocarcinoma of Mouse Prostate Model. Antioxid. Redox Signal. 2016, 24, 575–589. [Google Scholar] [CrossRef]
  138. Mandair, D.; Rossi, R.E.; Pericleous, M.; Whyand, T.; Caplin, M. Prostate cancer and the influence of dietary factors and supplements: A systematic review. Nutr. Metab. 2014, 11, 30. [Google Scholar] [CrossRef] [Green Version]
  139. Potosky, A.L.; Haque, R.; Cassidy-Bushrow, A.E.; Yood, M.U.; Jiang, M.; Tsai, H.-T.; Luta, G.; Keating, N.L.; Smith, M.R.; Eeden, S.K.V.D. Effectiveness of Primary Androgen-Deprivation Therapy for Clinically Localized Prostate Cancer. J. Clin. Oncol. 2014, 32, 1324–1330. [Google Scholar] [CrossRef] [Green Version]
  140. Du, F.; Sun, L.; Chu, Y.; Li, T.; Lei, C.; Wang, X.; Jiang, M.; Min, Y.; Lu, Y.; Zhao, X.; et al. DDIT4 promotes gastric cancer proliferation and tumorigenesis through the p53 and MAPK pathways. Cancer Commun. 2018, 38, 45. [Google Scholar] [CrossRef] [Green Version]
  141. Presti, D.; Quaquarini, E. The PI3K/AKT/mTOR and CDK4/6 Pathways in Endocrine Resistant HR+/HER2- Metastatic Breast Cancer: Biological Mechanisms and New Treatments. Cancers 2019, 11, 1242. [Google Scholar] [CrossRef] [Green Version]
  142. Dai, X.; Wang, L.; Deivasigamni, A.; Looi, C.Y.; Karthikeyan, C.; Trivedi, P.; Chinnathambi, A.; Alharbi, S.A.; Arfuso, F.; Dharmarajan, A.; et al. A novel benzimidazole derivative, MBIC inhibits tumor growth and promotes apoptosis via activation of ROS-dependent JNK signaling pathway in hepatocellular carcinoma. Oncotarget 2017, 8, 12831–12842. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Ong, P.S.; Wang, L.Z.; Dai, X.; Tseng, S.H.; Loo, S.J.; Sethi, G. Judicious Toggling of mTOR Activity to Combat Insulin Resistance and Cancer: Current Evidence and Perspectives. Front. Pharmacol. 2016, 7, 395. [Google Scholar] [CrossRef] [PubMed]
  144. Ryu, S.; Lim, W.; Bazer, F.W.; Song, G. Chrysin induces death of prostate cancer cells by inducing ROS and ER stress. J. Cell. Physiol. 2017, 232, 3786–3797. [Google Scholar] [CrossRef] [PubMed]
  145. Laudisi, F.; Di Grazia, A.; De Simone, V.; Cherubini, F.; Colantoni, A.; Ortenzi, A.; Franzè, E.; DiNallo, V.; Di Fusco, D.; Monteleone, I.; et al. Induction of endoplasmic reticulum stress and inhibition of colon carcinogenesis by the anti-helmintic drug rafoxanide. Cancer Lett. 2019, 462, 1–11. [Google Scholar] [CrossRef]
  146. Okubo, K.; Isono, M.; Asano, T.; Sato, A. Lopinavir-Ritonavir Combination Induces Endoplasmic Reticulum Stress and Kills Urological Cancer Cells. Anticancer Res. 2019, 39, 5891–5901. [Google Scholar] [CrossRef]
  147. Li, X.; Zeng, Z.; Wang, J.; Wu, Y.; Chen, W.; Zheng, L.; Xi, T.; Wang, A.; Lu, Y. MicroRNA-9 and breast cancer. Biomed. Pharmacother. 2019, 122, 109687. [Google Scholar] [CrossRef]
  148. Niyonizigiye, I.; Ngabire, D.; Patil, M.P.; Singh, A.A.; Kim, G.-D. In vitro induction of endoplasmic reticulum stress in human cervical adenocarcinoma HeLa cells by fucoidan. Int. J. Biol. Macromol. 2019, 137, 844–852. [Google Scholar] [CrossRef]
  149. Jessmon, P.; Boulanger, T.; Zhou, W.; Patwardhan, P. Epidemiology and treatment patterns of epithelial ovarian cancer. Expert Rev. Anticancer. Ther. 2017, 17, 427–437. [Google Scholar] [CrossRef]
  150. Guo, X.; Mei, J.; Jing, Y.; Wang, S. Curcumin-Loaded Nanoparticles with Low-Intensity Focused Ultrasound-Induced Phase Transformation as Tumor-Targeted and pH-Sensitive Theranostic Nanoplatform of Ovarian Cancer. Nanoscale Res. Lett. 2020, 15, 73. [Google Scholar] [CrossRef]
  151. Zhang, Y.; Yang, S.; Yang, Y.; Liu, T. Resveratrol induces immunogenic cell death of human and murine ovarian carcinoma cells. Infect. Agents Cancer 2019, 14, 27–29. [Google Scholar] [CrossRef] [Green Version]
  152. Bock, F.J.; Tait, S.W.G. Mitochondria as multifaceted regulators of cell death. Nat. Rev. Mol. Cell Biol. 2019, 21, 85–100. [Google Scholar] [CrossRef] [PubMed]
  153. Lim, W.; Ryu, S.; Bazer, F.W.; Kim, S.-M.; Song, G. Chrysin attenuates progression of ovarian cancer cells by regulating signaling cascades and mitochondrial dysfunction. J. Cell. Physiol. 2017, 233, 3129–3140. [Google Scholar] [CrossRef] [PubMed]
  154. Tewari, D.; Patni, P.; Bishayee, A.; Sah, A.N.; Bishayee, A. Natural products targeting the PI3K-Akt-mTOR signaling pathway in cancer: A novel therapeutic strategy. Semin. Cancer Biol. 2019. [Google Scholar] [CrossRef] [PubMed]
  155. Shi, N.; Yu, H.; Chen, T. Inhibition of esophageal cancer growth through the suppression of PI3K/AKT/mTOR signaling pathway. OncoTargets Ther. 2019, 12, 7637–7647. [Google Scholar] [CrossRef] [Green Version]
  156. Jiang, C.; Ma, Z.; Zhang, G.; Yang, X.; Du, Q.; Wang, W. CSNK2A1 Promotes Gastric Cancer Invasion Through the PI3K-Akt-mTOR Signaling Pathway. Cancer Manag. Res. 2019, 11, 10135–10143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Bray, F.; Ferlay, J.; Soerjomataram, I.; Siegel, R.L.; Torre, L.A.; Jemal, A. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2018, 68, 394–424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Manu, K.A.; Shanmugam, M.K.; Ramachandran, L.; Li, F.; Siveen, K.S.; Chinnathambi, A.; Zayed, M.; Alharbi, S.A.; Arfuso, F.; Kumar, A.P.; et al. Isorhamnetin augments the anti-tumor effect of capeciatbine through the negative regulation of NF-κB signaling cascade in gastric cancer. Cancer Lett. 2015, 363, 28–36. [Google Scholar] [CrossRef]
  159. Hwang, S.T.; Kim, C.; Lee, J.H.; Chinnathambi, A.; Alharbi, S.A.; Shair, O.H.; Sethi, G.; Ahn, K.S. Cycloastragenol can negate constitutive STAT3 activation and promote paclitaxel-induced apoptosis in human gastric cancer cells. Phytomedicine 2019, 59, 152907. [Google Scholar] [CrossRef]
  160. Manu, K.A.; Shanmugam, M.K.; Li, F.; Chen, L.; Siveen, K.S.; Ahn, K.S.; Kumar, A.P.; Sethi, G. Simvastatin sensitizes human gastric cancer xenograft in nude mice to capecitabine by suppressing nuclear factor-kappa B-regulated gene products. J. Mol. Med. 2013, 92, 267–276. [Google Scholar] [CrossRef] [Green Version]
  161. He, Y.-F.; Li, B.-Z.; Li, Z.; Liu, P.; Wang, Y.; Tang, Q.; Ding, J.; Jia, Y.; Chen, Z.; Li, L.; et al. Tet-Mediated Formation of 5-Carboxylcytosine and Its Excision by TDG in Mammalian DNA. Science 2011, 333, 1303–1307. [Google Scholar] [CrossRef] [Green Version]
  162. Yatagai, N.; Saito, T.; Akazawa, Y.; Hayashi, T.; Yanai, Y.; Tsuyama, S.; Ueyama, H.; Murakami, T.; Watanabe, S.; Nagahara, A.; et al. TP53 inactivation and expression of methylation-associated proteins in gastric adenocarcinoma with enteroblastic differentiation. Virchows Arch. 2018, 474, 315–324. [Google Scholar] [CrossRef] [PubMed]
  163. He, Z.; Wang, X.; Huang, C.; Gao, Y.; Yang, C.; Zeng, P.; Chen, Z. The FENDRR/miR-214-3P/TET2 axis affects cell malignant activity via RASSF1A methylation in gastric cancer. Am. J. Transl. Res. 2018, 10, 3211–3223. [Google Scholar] [PubMed]
  164. Du, C.; Kurabe, N.; Matsushima, Y.; Suzuki, M.; Kahyo, T.; Ohnishi, I.; Tanioka, F.; Tajima, S.; Goto, M.; Yamada, H.; et al. Robust quantitative assessments of cytosine modifications and changes in the expressions of related enzymes in gastric cancer. Gastric Cancer 2014, 18, 516–525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Zhong, X.; Liu, D.; Jiang, Z.; Li, C.; Chen, L.; Xia, Y.; Liu, D.; Yao, Q.; Wang, D. Chrysin Induced Cell Apoptosis and Inhibited Invasion Through Regulation of TET1 Expression in Gastric Cancer Cells. OncoTargets Ther. 2020, 13, 3277–3287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Han, Z.; Zhang, J.; Zhang, K.; Zhao, Y. Curcumin inhibits cell viability, migration, and invasion of thymic carcinoma cells via downregulation of microRNA-27a. Phytother. Res. 2020, 34, 1629–1637. [Google Scholar] [CrossRef] [PubMed]
  167. Zhu, M.; Zheng, Z.; Huang, J.; Ma, X.; Huang, C.; Wu, R.; Li, X.; Liang, Z.; Deng, F.; Wu, J.; et al. Modulation of miR-34a in curcumin-induced antiproliferation of prostate cancer cells. J. Cell. Biochem. 2019, 120, 15616–15624. [Google Scholar] [CrossRef]
  168. Boubaker, N.S.; Spagnuolo, M.; Trabelsi, N.; Said, R.; Gurtner, A.; Regazzo, G.; Ayed, H.; Blel, A.; Karray, O.; Saadi, A.; et al. miR-143 expression profiles in urinary bladder cancer: Correlation with clinical and epidemiological parameters. Mol. Biol. Rep. 2019, 47, 1283–1292. [Google Scholar] [CrossRef]
  169. Shirjang, S.; Mansoori, B.; Asghari, S.; Duijf, P.H.; Mohammadi, A.; Gjerstorff, M.F.; Baradaran, B. MicroRNAs in cancer cell death pathways: Apoptosis and necroptosis. Free Radic. Biol. Med. 2019, 139, 1–15. [Google Scholar] [CrossRef]
  170. Wang, J.; Zhang, L.; Jiang, W.; Zhang, R.; Zhang, B.; Silayiding, A.; Duan, X. MicroRNA-135a promotes proliferation, migration, invasion and induces chemoresistance of endometrial cancer cells. Eur. J. Obstet. Gynecol. Reprod. Biol. X 2019, 5, 100103. [Google Scholar] [CrossRef]
  171. Xu, C.; Li, B.; Zhao, S.; Jin, B.; Jia, R.; Ge, J.; Xu, H. MicroRNA-186-5p Inhibits Proliferation And Metastasis Of Esophageal Cancer By Mediating HOXA9. OncoTargets Ther. 2019, 12, 8905–8914. [Google Scholar] [CrossRef] [Green Version]
  172. Mohammadian, F.; Soltanahmadid, Y.P.; Zarghami, F.; Akbarzadeh, A.; Zarghami, N. Upregulation of miR-9 and Let-7a by nanoencapsulated chrysin in gastric cancer cells. Artif. Cells Nanomed. Biotechnol. 2016, 45, 1201–1206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Liu, H.; Wang, H.; Li, C.; Zhang, T.; Meng, X.; Zhang, Y.; Qian, H. Spheres from cervical cancer cells display stemness and cancer drug resistance. Oncol. Lett. 2016, 12, 2184–2188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Cheng, Y.-M.; Shen, C.-J.; Chang, C.-C.; Chou, C.-Y.; Tsai, C.-C.; Hsu, Y.-C. Inducement of apoptosis by cucurbitacin E, a tetracyclic triterpenes, through death receptor 5 in human cervical cancer cell lines. Cell Death Discov. 2017, 3, 17014. [Google Scholar] [CrossRef] [PubMed]
  175. Ningegowda, R.; Shivananju, N.S.; Rajendran, P.; Rangappa, K.S.; Chinnathambi, A.; Li, F.; Achar, R.R.; Shanmugam, M.K.; Bist, P.; Alharbi, S.A.; et al. A novel 4,6-disubstituted-1,2,4-triazolo-1,3,4-thiadiazole derivative inhibits tumor cell invasion and potentiates the apoptotic effect of TNFα by abrogating NF-κB activation cascade. Apoptosis 2016, 22, 145–157. [Google Scholar] [CrossRef]
  176. Rashid, S.; Labani, S.; Das, B.C. Knowledge, Awareness and Attitude on HPV, HPV Vaccine and Cervical Cancer among the College Students in India. PLoS ONE 2016, 11, e0166713. [Google Scholar] [CrossRef]
  177. Ortiz-Sánchez, E.; Santiago-López, L.; Cruz-Domínguez, V.B.; Toledo-Guzmán, M.E.; Hernández-Cueto, D.; Muñiz-Hernández, S.; Garrido, E.; De León, D.C.; García-Carrancá, A. Characterization of cervical cancer stem cell-like cells: Phenotyping, stemness, and human papilloma virus co-receptor expression. Oncotarget 2016, 7, 31943–31954. [Google Scholar] [CrossRef] [Green Version]
  178. Vishnoi, K.; Mahata, S.; Tyagi, A.; Pandey, A.; Verma, G.; Jadli, M.; Singh, T.; Singh, S.M.; Bharti, A.C. Cross-talk between Human Papillomavirus Oncoproteins and Hedgehog Signaling Synergistically Promotes Stemness in Cervical Cancer Cells. Sci. Rep. 2016, 6, 34377. [Google Scholar] [CrossRef] [Green Version]
  179. Hazelbag, S.; Fleuren, G.J.; Baelde, H.J.; Schuuring, E.; Kenter, G.G.; Gorter, A. Cytokine Profile of Cervical Cancer Cells. Gynecol. Oncol. 2001, 83, 235–243. [Google Scholar] [CrossRef]
  180. Wang, Q.; Zhang, C.; Walayat, S.; Chen, H.W.; Wang, Y. Association between cytokine gene polymorphisms and cervical cancer in a Chinese population. Eur. J. Obstet. Gynecol. Reprod. Biol. 2011, 158, 330–333. [Google Scholar] [CrossRef]
  181. Yang, B.; Lu, Y.; Zhang, A.; Zhou, A.; Zhang, L.; Zhang, L.; Gao, L.; Zang, Y.; Tang, X.; Sun, L. Doxycycline Induces Apoptosis and Inhibits Proliferation and Invasion of Human Cervical Carcinoma Stem Cells. PLoS ONE 2015, 10, e0129138. [Google Scholar]
  182. Dong, W.; Chen, A.; Chao, X.; Li, X.; Cui, Y.; Xu, C.; Cao, J.; Ning, Y.; Cao, X. Chrysin Inhibits Proinflammatory Factor-Induced EMT Phenotype and Cancer Stem Cell-Like Features in HeLa Cells by Blocking the NF-κB/Twist Axis. Cell. Physiol. Biochem. 2019, 52, 1236–1250. [Google Scholar] [CrossRef] [PubMed]
  183. Sinha, S.; Pokhrel, S.; Vaidya, B.N.; Joshee, N. In vitro micropropagation and callus induction in Scutellaria discolor colebr.–A medicinally important plant of Nepal. Indian J. Plant Genet. Resour. 1998, 12, 219–223. [Google Scholar]
  184. Laishram, S.; Moirangthem, D.S.; Borah, J.C.; Pal, B.C.; Suman, P.; Gupta, S.K.; Kalita, M.C.; Talukdar, N.C. Chrysin rich Scutellaria discolor Colebr. induces cervical cancer cell death via the induction of cell cycle arrest and caspase-dependent apoptosis. Life Sci. 2015, 143, 105–113. [Google Scholar] [CrossRef] [PubMed]
  185. Samec, M.; Líšková, A.; Koklesova, L.; Samuel, S.M.; Zhai, K.; Buhrmann, C.; Varghese, E.; Abotaleb, M.; Qaradakhi, T.; Zulli, A.; et al. Flavonoids against the Warburg phenotype—Concepts of predictive, preventive and personalised medicine to cut the Gordian knot of cancer cell metabolism. EPMA J. 2020, 11, 377–398. [Google Scholar] [CrossRef]
  186. Hanahan, D.; Weinberg, R.A. Hallmarks of Cancer: The Next Generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [Green Version]
  187. Heiden, M.G.V.; Cantley, L.C.; Thompson, C.B. Understanding the Warburg Effect: The Metabolic Requirements of Cell Proliferation. Science 2009, 324, 1029–1033. [Google Scholar] [CrossRef] [Green Version]
  188. Mathupala, S.P.; Ko, Y.H.; Pedersen, P.L. Hexokinase-2 bound to mitochondria: Cancer’s stygian link to the “Warburg effect” and a pivotal target for effective therapy. Semin. Cancer Biol. 2009, 19, 17–24. [Google Scholar] [CrossRef] [Green Version]
  189. Yoshida, G.J. Metabolic reprogramming: The emerging concept and associated therapeutic strategies. J. Exp. Clin. Cancer Res. 2015, 34, 1–10. [Google Scholar] [CrossRef] [Green Version]
  190. Rho, M.; Kim, J.; Jee, C.D.; Lee, Y.M.; Lee, H.E.; Kim, M.A.; Lee, H.S.; Kim, W.H. Expression of type 2 hexokinase and mitochondria-related genes in gastric carcinoma tissues and cell lines. Anticancer Res. 2007, 27, 251–258. [Google Scholar]
  191. Suh, D.H.; Kim, M.A.; Kim, H.; Kim, M.K.; Kim, H.-S.; Chung, H.H.; Kim, Y.B.; Song, Y.S. Association of overexpression of hexokinase II with chemoresistance in epithelial ovarian cancer. Clin. Exp. Med. 2013, 14, 345–353. [Google Scholar] [CrossRef]
  192. Palmieri, D.; Fitzgerald, D.; Shreeve, S.M.; Hua, E.; Bronder, J.L.; Weil, R.J.; Davis, S.; Stark, A.M.; Merino, M.J.; Kurek, R.; et al. Analyses of resected human brain metastases of breast cancer reveal the association between up-regulation of hexokinase 2 and poor prognosis. Mol. Cancer Res. 2009, 7, 1438–1445. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Guo-Qing, P.; Yuan, Y.; Cai-Gao, Z.; Hongling, Y.; Gonghua, H.; Yan, T. A study of association between expression of hOGG1, VDAC1, HK-2 and cervical carcinoma. J. Exp. Clin. Cancer Res. 2010, 29, 129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Xu, D.; Jin, J.; Yu, H.; Zhao, Z.; Ma, D.; Zhang, C.; Jiang, H. Chrysin inhibited tumor glycolysis and induced apoptosis in hepatocellular carcinoma by targeting hexokinase-2. J. Exp. Clin. Cancer Res. 2017, 36, 44. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Seydi, E.; Rahimpour, Z.; Salimi, A.; Pourahmad, J. Selective toxicity of chrysin on mitochondria isolated from liver of a HCC rat model. Bioorg. Med. Chem. 2019, 27, 115163. [Google Scholar] [CrossRef] [PubMed]
  196. Lee, C.; Cheung, S.T. STAT3: An Emerging Therapeutic Target for Hepatocellular Carcinoma. Cancers 2019, 11, 1646. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Swamy, S.G.; Kameshwar, V.H.; Shubha, P.B.; Looi, C.Y.; Shanmugam, M.K.; Arfuso, F.; Dharmarajan, A.; Sethi, G.; Swamy, S.N.; Bishayee, A. Targeting multiple oncogenic pathways for the treatment of hepatocellular carcinoma. Target. Oncol. 2016, 12, 1–10. [Google Scholar] [CrossRef]
  198. Dai, X.; Ahn, K.S.; Kim, C.; Siveen, K.S.; Ong, T.H.; Shanmugam, M.K.; Li, F.; Shi, J.; Kumar, A.P.; Wang, L.Z.; et al. Ascochlorin, an isoprenoid antibiotic inhibits growth and invasion of hepatocellular carcinoma by targeting STAT3 signaling cascade through the induction of PIAS3. Mol. Oncol. 2015, 9, 818–833. [Google Scholar] [CrossRef]
  199. Dai, X.; Ahn, K.S.; Wang, L.Z.; Kim, C.; Deivasigamni, A.; Arfuso, F.; Um, J.-Y.; Kumar, A.P.; Chang, Y.-C.; Kumar, D.; et al. Ascochlorin Enhances the Sensitivity of Doxorubicin Leading to the Reversal of Epithelial-to-Mesenchymal Transition in Hepatocellular Carcinoma. Mol. Cancer Ther. 2016, 15, 2966–2976. [Google Scholar] [CrossRef] [Green Version]
  200. Jiang, J.; Chen, Y.; Dong, T.; Yue, M.; Zhang, Y.; An, T.; Zhang, J.; Liu, P.; Yang, X. Polydatin inhibits hepatocellular carcinoma via the AKT/STAT3-FOXO1 signaling pathway. Oncol. Lett. 2019, 17, 4505–4513. [Google Scholar] [CrossRef] [Green Version]
  201. Mohan, C.D.; Bharathkumar, H.; Bulusu, K.C.; Pandey, V.; Rangappa, S.; Fuchs, J.E.; Shanmugam, M.K.; Dai, X.; Li, F.; Deivasigamani, A.; et al. Development of a novel azaspirane that targets the Janus kinase-signal transducer and activator of transcription (STAT) pathway in hepatocellular carcinoma in vitro and in vivo. J. Biol. Chem. 2014, 289, 34296–34307. [Google Scholar] [CrossRef] [Green Version]
  202. Mastron, J.K.; Siveen, K.S.; Sethi, G.; Bishayee, A. Silymarin and hepatocellular carcinoma. Anti-Cancer Drugs 2015, 26, 475–486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Zhang, Y.; Chen, F.; Xiao, X.; Pan, W.; Yuan, Q.; Cao, J. Chrysin inhibits sphere formation in SMMC-7721 cells via modulation of SHP-1/STAT3 signaling pathway. Cancer Manag. Res. 2019, 11, 2977–2985. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Kuphal, S.; Bosserhoff, A.-K. Recent progress in understanding the pathology of malignant melanoma. J. Pathol. 2009, 219, 400–409. [Google Scholar] [CrossRef] [PubMed]
  205. Morton, D.; Eilber, F.R.; Malmgren, R.A.; Wood, W.C. Immunological factors which influence response to immunotherapy in malignant melanoma. Surgery 1970, 68, 158. [Google Scholar]
  206. Bittner, M.; Meltzer, P.; Chen, Y.; Jiang, Y.; Seftor, E.; Hendrix, M.; Radmacher, M.; Simon, R.; Yakhini, Z.; Ben-Dor, A.; et al. Molecular classification of cutaneous malignant melanoma by gene expression profiling. Nature 2000, 406, 536–540. [Google Scholar] [CrossRef] [Green Version]
  207. Álvarez-Garcia, V.; Tawil, Y.; Wise, H.M.; Leslie, N.R. Mechanisms of PTEN loss in cancer: It’s all about diversity. Semin. Cancer Biol. 2019, 59, 66–79. [Google Scholar] [CrossRef]
  208. Ghazaeian, M.; Khorsandi, K.; Hosseinzadeh, R.; Naderi, A.; Abrahamse, H. Curcumin–silica nanocomplex preparation, hemoglobin and DNA interaction and photocytotoxicity against melanoma cancer cells. J. Biomol. Struct. Dyn. 2020, 1–11. [Google Scholar] [CrossRef]
  209. Yang, H.-Z.; Zhang, J.; Zeng, J.; Liu, S.; Zhou, F.; Zhang, F.; Giampieri, F.; Cianciosi, D.; Forbes-Hernandez, T.Y.; Ansary, J.; et al. Resveratrol inhibits the proliferation of melanoma cells by modulating cell cycle. Int. J. Food Sci. Nutr. 2019, 71, 84–93. [Google Scholar] [CrossRef]
  210. Chen, Y.; Li, H.; Zhang, G.-X.; Wu, Y.; Xiao, J.; Liu, J.; Qiu, P.; Liu, X.; Sun, L.; Du, B.; et al. Synergistic inhibitory effect of resveratrol and TK/GCV therapy on melanoma cells. J. Cancer Res. Clin. Oncol. 2020, 146, 1489–1499. [Google Scholar] [CrossRef]
  211. Sassi, A.; Maatouk, M.; El Gueder, D.; Bzéouich, I.M.; Hatira, S.A.-B.; Jemni-Yacoub, S.; Ghedira, K.; Chekir, L. Chrysin, a natural and biologically active flavonoid suppresses tumor growth of mouse B16F10 melanoma cells: In vitro and In vivo study. Chem. Interact. 2018, 283, 10–19. [Google Scholar] [CrossRef]
  212. Folgueras, A.R.; Pendas, A.M.; Sánchez, L.M.; López-Otín, C. Matrix metalloproteinases in cancer: From new functions to improved inhibition strategies. Int. J. Dev. Biol. 2004, 48, 411–424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Kessenbrock, K.; Plaks, V.; Werb, Z. Matrix metalloproteinases: Regulators of the tumor microenvironment. Cell 2010, 141, 52–67. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Babykutty, S.; Suboj, P.; Srinivas, P.; Nair, A.S.; Chandramohan, K.; Gopala, S. Insidious role of nitric oxide in migration/invasion of colon cancer cells by upregulating MMP-2/9 via activation of cGMP-PKG-ERK signaling pathways. Clin. Exp. Metastasis 2012, 29, 471–492. [Google Scholar] [CrossRef] [PubMed]
  215. Dung, T.D.; Feng, C.C.; Kuo, W.W.; Pai, P.; Chung, L.C.; Chang, S.H.; Hsu, H.H.; Tsai, F.J.; Lin, Y.M.; Huang, C.Y. Suppression of plasminogen activators and the MMP-2/-9 pathway by a Zanthoxylum avicennae extract to inhibit the HA22T human hepatocellular carcinoma cell migration and invasion effects in vitro and in vivo via phosphatase 2A activation. Biosci. Biotechnol. Biochem. 2013, 77, 1814–1821. [Google Scholar] [CrossRef]
  216. Cai, Z.; Zhang, F.; Chen, W.; Zhang, J.; Li, H. miRNAs: A Promising Target in the Chemoresistance of Bladder Cancer. OncoTargets Ther. 2019, 12, 11805–11816. [Google Scholar] [CrossRef] [Green Version]
  217. Xu, R.; Li, H.; Wu, S.; Qu, J.; Yuan, H.; Zhou, Y.; Lu, Q. MicroRNA-1246 regulates the radio-sensitizing effect of curcumin in bladder cancer cells via activating P53. Int. Urol. Nephrol. 2019, 51, 1771–1779. [Google Scholar] [CrossRef]
  218. Almeida, T.C.; Guerra, C.C.C.; De Assis, B.L.G.; Soares, R.D.O.A.; Garcia, C.C.M.; Lima, A.A.; Da Silva, G.N. Antiproliferative and toxicogenomic effects of resveratrol in bladder cancer cells with different TP53 status. Environ. Mol. Mutagen. 2019, 60, 740–751. [Google Scholar] [CrossRef]
  219. Sun, N.; Liang, Y.; Chen, Y.; Wang, L.; Li, D.; Liang, Z.; Sun, L.; Wang, Y.; Niu, H. Glutamine affects T24 bladder cancer cell proliferation by activating STAT3 through ROS and glutaminolysis. Int. J. Mol. Med. 2019, 44, 2189–2200. [Google Scholar] [CrossRef]
  220. Korac-Prlic, J.; Degoricija, M.; Vilović, K.; Haupt, B.; Ivanišević, T.; Franković, L.; Grivennikov, S.; Terzić, J. Targeting Stat3 signaling impairs the progression of bladder cancer in a mouse model. Cancer Lett. 2020, 490, 89–99. [Google Scholar] [CrossRef]
  221. Anand, V.; Khandelwal, M.; Appunni, S.; Gupta, N.; Seth, A.; Singh, P.; Mathur, S.; Sharma, A. CD44 splice variant (CD44v3) promotes progression of urothelial carcinoma of bladder through Akt/ERK/STAT3 pathways: Novel therapeutic approach. J. Cancer Res. Clin. Oncol. 2019, 145, 2649–2661. [Google Scholar] [CrossRef]
  222. Xu, Y.; Tong, Y.; Ying, J.; Lei, Z.; Wan, L.; Zhu, X.; Ye, F.; Mao, P.; Wu, X.; Pan, R.; et al. Chrysin induces cell growth arrest, apoptosis, and ER stress and inhibits the activation of STAT3 through the generation of ROS in bladder cancer cells. Oncol. Lett. 2018, 15, 9117–9125. [Google Scholar] [CrossRef] [PubMed]
  223. Puccini, A.; Berger, M.D.; Naseem, M.; Tokunaga, R.; Battaglin, F.; Cao, S.; Hanna, D.L.; McSkane, M.; Soni, S.; Zhang, W.; et al. Colorectal cancer: Epigenetic alterations and their clinical implications. Biochim. Biophys. Acta (BBA) Bioenergy 2017, 1868, 439–448. [Google Scholar] [CrossRef] [PubMed]
  224. Coppedè, F.; Lopomo, A.; Spisni, R.; Migliore, L. Genetic and epigenetic biomarkers for diagnosis, prognosis and treatment of colorectal cancer. World J. Gastroenterol. 2014, 20, 943–956. [Google Scholar] [CrossRef] [PubMed]
  225. Lee, C.S.; Ryan, E.J.; Doherty, G.A. Gastro-intestinal toxicity of chemotherapeutics in colorectal cancer: The role of inflammation. World J. Gastroenterol. 2014, 20, 3751–3761. [Google Scholar] [CrossRef]
  226. Crea, F.; Nobili, S.; Paolicchi, E.; Perrone, G.; Napoli, C.; Landini, I.; Danesi, R.; Mini, E. Epigenetics and chemoresistance in colorectal cancer: An opportunity for treatment tailoring and novel therapeutic strategies. Drug Resist. Updat. 2011, 14, 280–296. [Google Scholar] [CrossRef]
  227. Lin, Y.-M.; Chen, C.-I.; Hsiang, Y.-P.; Hsu, Y.-C.; Cheng, K.-C.; Chien, P.-H.; Pan, H.-L.; Lu, C.-C.; Chen, Y.-J. Chrysin Attenuates Cell Viability of Human Colorectal Cancer Cells through Autophagy Induction Unlike 5-Fluorouracil/Oxaliplatin. Int. J. Mol. Sci. 2018, 19, 1763. [Google Scholar] [CrossRef] [Green Version]
  228. Eghtedardoost, M.; Ghazanfari, T.; Sadeghipour, A.; Hassan, Z.M.; Ghanei, M.; Ghavami, S. Delayed effects of sulfur mustard on autophagy suppression in chemically-injured lung tissue. Int. Immunopharmacol. 2020, 80, 105896. [Google Scholar] [CrossRef]
  229. Patra, S.; Mishra, S.R.; Behera, B.P.; Mahapatra, K.K.; Panigrahi, D.P.; Bhol, C.S.; Praharaj, P.P.; Sethi, G.; Patra, S.K.; Bhutia, S.K. Autophagy-modulating phytochemicals in cancer therapeutics: Current evidences and future perspectives. Semin. Cancer Biol. 2020. [Google Scholar] [CrossRef]
  230. Linder, B.; Kögel, D. Autophagy in Cancer Cell Death. Biology 2019, 8, 82. [Google Scholar] [CrossRef] [Green Version]
  231. Teng, J.-F.; Qin, D.-L.; Mei, Q.-B.; Qiu, W.-Q.; Pan, R.; Xiong, R.; Zhao, Y.; Law, B.Y.-K.; Wong, V.K.-W.; Tang, Y.; et al. Polyphyllin VI, a saponin from Trillium tschonoskii Maxim. induces apoptotic and autophagic cell death via the ROS triggered mTOR signaling pathway in non-small cell lung cancer. Pharmacol. Res. 2019, 147, 104396. [Google Scholar] [CrossRef]
  232. Majeed, T.; Wani, I.A.; Hamdani, A.M.; Bhat, N.A. Effect of sonication and γ-irradiation on the properties of pea (Pisum sativum) and vetch (Vicia villosa) starches: A comparative study. Int. J. Biol. Macromol. 2018, 114, 1144–1150. [Google Scholar] [CrossRef] [PubMed]
  233. Jenjob, A.; Uthairatanakij, A.; Jitareerat, P.; Wongs-Aree, C.; Aiamla-Or, S. Effect of harvest seasonal and gamma irradiation on the physicochemical changes in pineapple fruit cv. Pattavia during stimulated sea shipment. Food Sci. Nutr. 2017, 5, 997–1003. [Google Scholar] [CrossRef] [PubMed]
  234. Kang, J.A.; Song, H.-Y.; Byun, E.-H.; Ahn, N.-G.; Kim, H.-M.; Nam, Y.R.; Lee, G.H.; Jang, B.-S.; Choi, D.S.; Lee, D.-E.; et al. Gamma-irradiated black ginseng extract inhibits mast cell degranulation and suppresses atopic dermatitis-like skin lesions in mice. Food Chem. Toxicol. 2018, 111, 133–143. [Google Scholar] [CrossRef] [PubMed]
  235. Song, H.-Y.; Kim, H.-M.; Mushtaq, S.; Kim, W.S.; Kim, Y.J.; Lim, S.-T.; Byun, E.-B. Gamma-Irradiated Chrysin Improves Anticancer Activity in HT-29 Colon Cancer Cells Through Mitochondria-Related Pathway. J. Med. Food 2019, 22, 713–721. [Google Scholar] [CrossRef] [PubMed]
  236. Chinetti-Gbaguidi, G.; Fruchart, J.-C.; Staels, B. Peroxisome proliferator-activated receptors (PPARs): Nuclear receptors at the crossroads between lipid metabolism and inflammation. Inflamm. Res. 2000, 49, 497–505. [Google Scholar] [CrossRef] [PubMed]
  237. Weiskirchen, S.; Weiskirchen, R. Resveratrol: How Much Wine Do You Have to Drink to Stay Healthy? Adv. Nutr. 2016, 7, 706–718. [Google Scholar] [CrossRef] [Green Version]
  238. Sikka, S.; Chen, L.; Sethi, G.; Kumar, A.P. Targeting PPARγ Signaling Cascade for the Prevention and Treatment of Prostate Cancer. PPAR Res. 2012, 2012, 1–14. [Google Scholar] [CrossRef] [Green Version]
  239. Ramachandran, L.; Manu, K.A.; Shanmugam, M.K.; Li, F.; Siveen, K.S.; Vali, S.; Kapoor, S.; Abbasi, T.; Surana, R.; Smoot, D.T.; et al. Isorhamnetin Inhibits Proliferation and Invasion and Induces Apoptosis through the Modulation of Peroxisome Proliferator-activated Receptor γ Activation Pathway in Gastric Cancer. J. Biol. Chem. 2012, 287, 38028–38040. [Google Scholar] [CrossRef] [Green Version]
  240. Berthold, D.R.; Pond, G.R.; Roessner, M.; De Wit, R.; Eisenberger, M.; Tannock, A.I.F. Treatment of Hormone-Refractory Prostate Cancer with Docetaxel or Mitoxantrone: Relationships between Prostate-Specific Antigen, Pain, and Quality of Life Response and Survival in the TAX-327 Study. Clin. Cancer Res. 2008, 14, 2763–2767. [Google Scholar] [CrossRef] [Green Version]
  241. Grau, R.; Punzón, C.; Fresno, M.; Iñiguez, M.A. Peroxisome-proliferator-activated receptor α agonists inhibit cyclo-oxygenase 2 and vascular endothelial growth factor transcriptional activation in human colorectal carcinoma cells via inhibition of activator protein-1. Biochem. J. 2006, 395, 81–88. [Google Scholar] [CrossRef] [Green Version]
  242. Waxman, D.J. P450 Gene Induction by Structurally Diverse Xenochemicals: Central Role of Nuclear Receptors CAR, PXR, and PPAR. Arch. Biochem. Biophys. 1999, 369, 11–23. [Google Scholar] [CrossRef] [PubMed]
  243. Danielson, P.á. The Cytochrome P450 Superfamily: Biochemistry, Evolution and Drug Metabolism in Humans. Curr. Drug Metab. 2002, 3, 561–597. [Google Scholar] [CrossRef] [PubMed]
  244. Thomas, M.; Winter, S.; Klumpp, B.; Turpeinen, M.; Klein, K.; Schwab, M.; Zanger, U.M. Peroxisome proliferator-activated receptor alpha, PPARα, directly regulates transcription of cytochrome P450 CYP2C8. Front. Pharmacol. 2015, 6, 261. [Google Scholar] [CrossRef] [Green Version]
  245. Khor, C.Y.; Khoo, B.Y. PPARα plays an important role in the migration activity, and the expression of CYP2S1 and CYP1B1 in chrysin-treated HCT116 cells. Biotechnol. Lett. 2020, 42, 1–15. [Google Scholar] [CrossRef] [PubMed]
  246. Tang, Q.; Ji, F.; Guo, J.; Wang, J.; Li, Y.; Bao, Y. Directional modification of chrysin for exerting apoptosis and enhancing significantly anti-cancer effects of 10-hydroxy camptothecin. Biomed. Pharmacother. 2016, 82, 693–703. [Google Scholar] [CrossRef] [PubMed]
  247. Li, H.Z.; Chen, Y.H.; Fang, Y.L.; Zhong, L.Y.; Yuan, Q.Q.; Xu, X.Y.; Cao, J.G. Effects of chrysin on sphere formation and CK2α expression of ovarian cancer stem-like cells derived from SKOV3 cell line. Zhonghua Yi Xue Za Zhi 2016, 96, 2013–2016. [Google Scholar] [PubMed]
  248. Samarghandian, S.; Afshari, J.T.; Davoodi, S. Chrysin reduces proliferation and induces apoptosis in the human prostate cancer cell line pc-3. Clinics 2011, 66, 1073–1079. [Google Scholar] [CrossRef] [Green Version]
  249. Al-Oudat, B.A.; Alqudah, M.A.; Audat, S.A.; Al-Balas, Q.A.; El-Elimat, T.; Hassan, M.A.; Frhat, I.N.; Azaizeh, M.M. Design, synthesis, and biologic evaluation of novel chrysin derivatives as cytotoxic agents and caspase-3/7 activators. Drug Des. Devel. Ther. 2019, 13, 423–433. [Google Scholar] [CrossRef] [Green Version]
  250. Xie, Y.; Peng, X. Effects of chrysin on the apoptosis in oral squamous carcinoma KB cell line and the underlying mechanisms. Zhong Nan Da Xue Xue Bao Yi Xue Ban 2019, 44, 522–527. [Google Scholar]
  251. Ni, Z.; He, J.; Wu, Y.; Hu, C.; Dai, X.; Yan, X.; Li, B.; Li, X.; Xiong, H.; Li, Y.; et al. AKT-mediated phosphorylation of ATG4B impairs mitochondrial activity and enhances the Warburg effect in hepatocellular carcinoma cells. Autophagy 2018, 14, 685–701. [Google Scholar] [CrossRef] [Green Version]
  252. Hasanzadeh, M.; Samarghandian, S.; Azimi-Nezhad, M.; Borji, A.; Jabbari, F.; Farkhondeh, T.; Samini, M. Inhibitory and cytotoxic activities of chrysin on human breast adenocarcinoma cells by induction of apoptosis. Pharmacogn. Mag. 2016, 12, 436–S440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Ren, S.-Z.; Wang, Z.-C.; Zhu, X.-H.; Zhu, D.; Li, Z.; Shen, F.-Q.; Duan, Y.-T.; Cao, H.; Zhao, J.; Zhu, H.-L. Design and biological evaluation of novel hybrids of 1, 5-diarylpyrazole and Chrysin for selective COX-2 inhibition. Bioorg. Med. Chem. 2018, 26, 4264–4275. [Google Scholar] [CrossRef] [PubMed]
  254. Yu, X.-M.; Phan, T.; Patel, P.N.; Jaskula-Sztul, R.; Chen, H.; Bs, P.N.P. Chrysin activates Notch1 signaling and suppresses tumor growth of anaplastic thyroid carcinoma in vitro and in vivo. Cancer 2012, 119, 774–781. [Google Scholar] [CrossRef] [Green Version]
  255. Nagasaka, M.; Hashimoto, R.; Inoue, Y.; Ishiuchi, K.; Matsuno, M.; Itoh, Y.; Tokugawa, M.; Ohoka, N.; Morishita, D.; Mizukami, H.; et al. Anti-Tumorigenic Activity of Chrysin from Oroxylum indicum via Non-Genotoxic p53 Activation through the ATM-Chk2 Pathway. Molecules 2018, 23, 1394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Celinska-Janowicz, K.; Zareba, I.; Lazarek, U.; Teul, J.; Tomczyk, M.; Pałka, J.; Miltyk, W. Constituents of Propolis: Chrysin, Caffeic Acid, p-Coumaric Acid, and Ferulic Acid Induce PRODH/POX-Dependent Apoptosis in Human Tongue Squamous Cell Carcinoma Cell (CAL-27). Front. Pharmacol. 2018, 9, 336. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Park, W.; Park, S.; Lim, W.; Song, G. Chrysin disrupts intracellular homeostasis through mitochondria-mediated cell death in human choriocarcinoma cells. Biochem. Biophys. Res. Commun. 2018, 503, 3155–3161. [Google Scholar] [CrossRef] [PubMed]
  258. Bahadori, M.; Baharara, J.; Amini, E. Anticancer Properties of Chrysin on Colon Cancer Cells, In vitro and In vivo with Modulation of Caspase-3, -9, Bax and Sall4. Iran. J. Biotechnol. 2016, 14, 177–184. [Google Scholar] [CrossRef] [Green Version]
  259. Kahroba, H.; Shirmohamadi, M.; Hejazi, M.S.; Samadi, N. The Role of Nrf2 signaling in cancer stem cells: From stemness and self-renewal to tumorigenesis and chemoresistance. Life Sci. 2019, 239, 116986. [Google Scholar] [CrossRef]
  260. Liu, M.-L.; Zang, F.; Zhang, S.-J. RBCK1 contributes to chemoresistance and stemness in colorectal cancer (CRC). Biomed. Pharmacother. 2019, 118, 109250. [Google Scholar] [CrossRef]
  261. Kirtonia, A.; Gala, K.; Fernandes, S.G.; Pandya, G.; Pandey, A.K.; Sethi, G.; Khattar, E.; Garg, M. Repurposing of drugs: An attractive pharmacological strategy for cancer therapeutics. Semin. Cancer Biol. 2020. [Google Scholar] [CrossRef]
  262. Paciello, F.; Fetoni, A.R.; Mezzogori, D.; Rolesi, R.; Di Pino, A.; Paludetti, G.; Grassi, C.; Troiani, D. The dual role of curcumin and ferulic acid in counteracting chemoresistance and cisplatin-induced ototoxicity. Sci. Rep. 2020, 10, 1–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. San, T.T.; Khaenam, P.; Prachayasittikul, V.; Sripa, B.; Kunkeaw, N.; Chan-On, W. Curcumin enhances chemotherapeutic effects and suppresses ANGPTL4 in anoikis-resistant cholangiocarcinoma cells. Heliyon 2020, 6, e03255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Soni, V.K.; Shukla, D.; Kumar, A.; Vishvakarma, N.K. Curcumin circumvent lactate-induced chemoresistance in hepatic cancer cells through modulation of hydroxycarboxylic acid receptor-1. Int. J. Biochem. Cell Biol. 2020, 123, 105752. [Google Scholar] [CrossRef] [PubMed]
  265. Lu, X.; Yang, F.; Chen, D.; Zhao, Q.; Chen, D.; Ping, H.; Xing, N. Quercetin reverses docetaxel resistance in prostate cancer via androgen receptor and PI3K/Akt signaling pathways. Int. J. Biol. Sci. 2020, 16, 1121–1134. [Google Scholar] [CrossRef]
  266. Xu, X.; Jiang, X.; Chen, L.; Zhao, Y.; Huang, Z.; Zhou, H.; Shi, M. MiR-181a Promotes Apoptosis and Reduces Cisplatin Resistance by Inhibiting Osteopontin in Cervical Cancer Cells. Cancer Biother. Radiopharm. 2019, 34, 559–565. [Google Scholar] [CrossRef]
  267. Zou, A.; Wu, A.; Luo, M.; Zhou, C.; Lu, Y.; Yu, X. SHCBP1 promotes cisplatin induced apoptosis resistance, migration and invasion through activating Wnt pathway. Life Sci. 2019, 235, 116798. [Google Scholar] [CrossRef]
  268. Liu, Y.; Gu, S.; Li, H.; Wang, J.; Wei, C.; Liu, Q. SNHG16 promotes osteosarcoma progression and enhances cisplatin resistance by sponging miR-16 to upregulate ATG4B expression. Biochem. Biophys. Res. Commun. 2019, 518, 127–133. [Google Scholar] [CrossRef]
  269. Tan, L.-M.; Li, X.; Qiu, C.-F.; Zhu, T.; Hu, C.-P.; Yin, J.-Y.; Zhang, W.; Zhou, H.-H.; Liu, Z.-Q. CLEC4M is associated with poor prognosis and promotes cisplatin resistance in NSCLC patients. J. Cancer 2019, 10, 6374–6383. [Google Scholar] [CrossRef] [Green Version]
  270. Martins, I.L.; Charneira, C.; Gandin, V.; Silva, J.L.A.F.; Justino, G.C.; Telo, J.P.; Vieira, A.J.S.C.; Marzano, C.; Antunes, A.M.M. Selenium-Containing Chrysin and Quercetin Derivatives: Attractive Scaffolds for Cancer Therapy. J. Med. Chem. 2015, 58, 4250–4265. [Google Scholar] [CrossRef]
  271. Kodumudi, K.N.; Woan, K.; Gilvary, D.L.; Sahakian, E.; Wei, S.; Djeu, J. A novel chemoimmunomodulating property of docetaxel: Suppression of myeloid-derived suppressor cells in tumor bearers. Clin. Cancer Res. 2010, 16, 4583–4594. [Google Scholar] [CrossRef] [Green Version]
  272. Sekino, Y.; Han, X.; Kawaguchi, T.; Babasaki, T.; Goto, K.; Inoue, S.; Hayashi, T.; Teishima, J.; Shiota, M.; Yasui, W.; et al. TUBB3 Reverses Resistance to Docetaxel and Cabazitaxel in Prostate Cancer. Int. J. Mol. Sci. 2019, 20, 3936. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  273. Tanaudommongkon, I.; Tanaudommongkon, A.; Prathipati, P.; Nguyen, J.T.; Keller, E.T.; Dong, X. Curcumin Nanoparticles and Their Cytotoxicity in Docetaxel-Resistant Castration-Resistant Prostate Cancer Cells. Biomedicines 2020, 8, 253. [Google Scholar] [CrossRef] [PubMed]
  274. Lim, H.-K.; Kim, K.M.; Jeong, S.-Y.; Choi, E.K.; Jung, J. Chrysin Increases the Therapeutic Efficacy of Docetaxel and Mitigates Docetaxel-Induced Edema. Integr. Cancer Ther. 2016, 16, 496–504. [Google Scholar] [CrossRef] [PubMed]
  275. Siddik, Z.H. Cisplatin: Mode of cytotoxic action and molecular basis of resistance. Oncogene 2003, 22, 7265–7279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  276. Aylon, Y.; Oren, M. Living with p53, Dying of p53. Cell 2007, 130, 597–600. [Google Scholar] [CrossRef] [Green Version]
  277. Nag, S.A.; Qin, J.-J.; Srivenugopal, K.S.; Wang, M.; Zhang, R. The MDM2-p53 pathway revisited. J. Biomed. Res. 2013, 27, 254–271. [Google Scholar] [CrossRef]
  278. Moll, U.M.; Wolff, S.; Speidel, D.; Deppert, W. Transcription-independent pro-apoptotic functions of p53. Curr. Opin. Cell Biol. 2005, 17, 631–636. [Google Scholar] [CrossRef]
  279. Li, X.; Huang, J.-M.; Wang, J.-N.; Xiong, X.-K.; Yang, X.-F.; Zou, F. Combination of chrysin and cisplatin promotes the apoptosis of Hep G2 cells by up-regulating p53. Chem. Interact. 2015, 232, 12–20. [Google Scholar] [CrossRef]
  280. Bhattacharjee, S.; Li, J.; Dashwood, R. Emerging crosstalk between long non-coding RNAs and Nrf2 signaling. Cancer Lett. 2020. [Google Scholar] [CrossRef]
  281. Li, L.; Chen, Y.; Jiao, D.; Yang, S.; Li, P. Protective Effect of Astaxanthin on Ochratoxin A-Induced Kidney Injury to Mice by Regulating Oxidative Stress-Related NRF2/KEAP1 Pathway. Molecules 2020, 25, 1386. [Google Scholar] [CrossRef] [Green Version]
  282. Wang, X.; Chang, X.; Zhan, H.; Zhang, Q.; Li, C.; Gao, Q.; Yang, M.; Luo, Z.; Li, S.; Sun, Y. Curcumin and Baicalin ameliorate ethanol-induced liver oxidative damage via the Nrf2/HO-1 pathway. J. Food Biochem. 2020, e13425. [Google Scholar] [CrossRef]
  283. Ji, L.; Zhang, R.; Chen, J.; Xue, Q.; Moghal, N.; Tsao, M.-S. PIDD interaction with KEAP1 as a new mutation-independent mechanism to promote NRF2 stabilization and chemoresistance in NSCLC. Sci. Rep. 2019, 9, 12437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Wu, J.; Zhang, L.; Li, H.; Wu, S.; Liu, Z. Nrf2 induced cisplatin resistance in ovarian cancer by promoting CD99 expression. Biochem. Biophys. Res. Commun. 2019, 518, 698–705. [Google Scholar] [CrossRef]
  285. Xia, X.; Wang, Q.; Ye, T.; Liu, Y.; Liu, D.; Song, S.; Zheng, C. NRF2/ABCB1-mediated efflux and PARP1-mediated dampening of DNA damage contribute to doxorubicin resistance in chronic hypoxic HepG2 cells. Fundam. Clin. Pharmacol. 2020, 34, 41–50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  286. Cucci, M.A.; Grattarola, M.; Dianzani, C.; Damia, G.; Ricci, F.; Roetto, A.; Trotta, F.; Barrera, G.; Pizzimenti, S. Ailanthone increases oxidative stress in CDDP-resistant ovarian and bladder cancer cells by inhibiting of Nrf2 and YAP expression through a post-translational mechanism. Free Radic. Biol. Med. 2020, 150, 125–135. [Google Scholar] [CrossRef]
  287. Gao, A.M.; Ke, Z.P.; Shi, F.; Sun, G.C.; Chen, H. Chrysin enhances sensitivity of BEL-7402/ADM cells to doxorubicin by suppressing PI3K/Akt/Nrf2 and ERK/Nrf2 pathway. Chem. Biol. Interact. 2013, 206, 100–108. [Google Scholar] [CrossRef]
  288. Laredj-Bourezg, F.; Bolzinger-Thevenin, M.A.; Pelletier, J.; Chevalier, Y. Pickering emulsions stabilized by biodegradable block copolymer micelles for controlled topical drug delivery. Int. J. Pharm. 2017, 531, 134–142. [Google Scholar] [CrossRef]
  289. Mandal, A.; Bisht, R.; Rupenthal, I.D.; Mitra, A.K. Polymeric micelles for ocular drug delivery: From structural frameworks to recent preclinical studies. J. Control. Release 2017, 248, 96–116. [Google Scholar] [CrossRef] [Green Version]
  290. Cabral, H.; Miyata, K.; Osada, K.; Kataoka, K. Block Copolymer Micelles in Nanomedicine Applications. Chem. Rev. 2018, 118, 6844–6892. [Google Scholar] [CrossRef] [Green Version]
  291. Arranja, A.G.; Pathak, V.; Lammers, T.; Shi, Y. Tumor-targeted nanomedicines for cancer theranostics. Pharmacol. Res. 2017, 115, 87–95. [Google Scholar] [CrossRef] [Green Version]
  292. Ghamkhari, A.; Pouyafar, A.; Salehi, R.; Rahbarghazi, R. Chrysin and Docetaxel Loaded Biodegradable Micelle for Combination Chemotherapy of Cancer Stem Cell. Pharm. Res. 2019, 36, 165. [Google Scholar] [CrossRef] [PubMed]
  293. Davaran, S.; Fazeli, H.; Ghamkhari, A.; Rahimi, F.; Molavi, O.; Anzabi, M.; Salehi, R. Synthesis and characterization of novel P(HEMA-LA-MADQUAT) micelles for co-delivery of methotrexate and Chrysin in combination cancer chemotherapy. J. Biomater. Sci. Polym. Ed. 2018, 29, 1265–1286. [Google Scholar] [CrossRef] [PubMed]
  294. Restani, R.B.; Morgado, P.I.; Ribeiro, M.P.; Correia, I.J.; Aguiar-Ricardo, A.; Bonifácio, V.D.B. Biocompatible Polyurea Dendrimers with pH-Dependent Fluorescence. Angew. Chem. 2012, 124, 5252–5255. [Google Scholar] [CrossRef]
  295. Palakurthi, S.; Yellepeddi, V.; Vangara, K.K. Recent trends in cancer drug resistance reversal strategies using nanoparticles. Expert Opin. Drug Deliv. 2012, 9, 287–301. [Google Scholar] [CrossRef]
  296. Brannon-Peppas, L.; Blanchette, J.O. Nanoparticle and targeted systems for cancer therapy. Adv. Drug Deliv. Rev. 2012, 64, 206–212. [Google Scholar] [CrossRef]
  297. Santos, I.; Ramos, C.; Mendes, C.; Sequeira, C.O.; Tomé, C.S.; Fernandes, D.G.; Mota, P.; Pires, R.F.; Urso, D.; Hipólito, A.; et al. Targeting Glutathione and Cystathionine β-Synthase in Ovarian Cancer Treatment by Selenium-Chrysin Polyurea Dendrimer Nanoformulation. Nutrients 2019, 11, 2523. [Google Scholar] [CrossRef] [Green Version]
  298. Eatemadi, A.; Darabi, M.; Afraidooni, L.; Zarghami, N.; Daraee, H.; Eskandari, L.; Mellatyar, H.; Akbarzadeh, A. Comparison, synthesis and evaluation of anticancer drug-loaded polymeric nanoparticles on breast cancer cell lines. Artif. Cells Nanomed. Biotechnol. 2015, 44, 1–10. [Google Scholar] [CrossRef]
  299. Seidi, K. Nanomagnet-Based Detoxifying Machine: An Alternative/Complementary Approach in HIV therapy. J. AIDS Clin. Res. 2014, 5. [Google Scholar] [CrossRef] [Green Version]
  300. Eatemadi, A.; Daraee, H.; Karimkhanloo, H.; Kouhi, M.; Zarghami, N.; Akbarzadeh, A.; Abasi, M.; Hanifehpour, Y.; Joo, S.W. Carbon nanotubes: Properties, synthesis, purification, and medical applications. Nanoscale Res. Lett. 2014, 9, 393. [Google Scholar] [CrossRef] [Green Version]
  301. Daraee, H.; Eatemadi, A.; Abbasi, E.; Aval, S.F.; Kouhi, M.; Akbarzadeh, A. Application of gold nanoparticles in biomedical and drug delivery. Artif. Cells Nanomed. Biotechnol. 2014, 44, 410–422. [Google Scholar] [CrossRef]
  302. Daraee, H.; Eatemadi, A.; Kouhi, M.; Alimirzalu, S.; Akbarzadeh, A. Application of liposomes in medicine and drug delivery. Artif. Cells Nanomed. Biotechnol. 2014, 44, 1–11. [Google Scholar] [CrossRef] [PubMed]
  303. Mellatyar, H.; Akbarzadeh, A.; Rahmati, M.; Ghalhar, M.G.; Eatemadi, A.; Nejati-Koshki, K.; Zarghami, N.; Barkhordari, A. Comparison of Inhibitory Effect of 17-DMAG Nanoparticles and Free 17-DMAG in HSP90 Gene Expression in Lung Cancer. Asian Pac. J. Cancer Prev. 2014, 15, 8693–8698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Eatemadi, A.; Daraee, H.; Aiyelabegan, H.T.; Negahdari, B.; Rajeian, B.; Zarghami, N. Synthesis and Characterization of Chrysin-loaded PCL-PEG-PCL nanoparticle and its effect on breast cancer cell line. Biomed. Pharmacother. 2016, 84, 1915–1922. [Google Scholar] [CrossRef] [PubMed]
  305. DiMarco-Crook, C.; Rakariyatham, K.; Li, Z.; Du, Z.; Zheng, J.; Wu, X.; Xiao, H. Synergistic anticancer effects of curcumin and 3’,4’-didemethylnobiletin in combination on colon cancer cells. J. Food Sci. 2020, 85, 1292–1301. [Google Scholar] [CrossRef]
  306. Verma, A.H.; Kumar, T.S.S.; Madhumathi, K.; Rubaiya, Y.; Ramalingan, M.; Doble, M. Curcumin Releasing Eggshell Derived Carbonated Apatite Nanocarriers for Combined Anti-Cancer, Anti-Inflammatory and Bone Regenerative Therapy. J. Nanosci. Nanotechnol. 2019, 19, 6872–6880. [Google Scholar] [CrossRef] [PubMed]
  307. Bagheri, R.; Sanaat, Z.; Zarghami, N. Synergistic Effect of Free and Nano-encapsulated Chrysin-Curcumin on Inhibition of hTERT Gene Expression in SW480 Colorectal Cancer Cell Line. Drug Res. 2018, 68, 335–343. [Google Scholar] [CrossRef]
  308. Kazemi-Lomedasht, F.; Rami, A.; Zarghami, N. Comparison of Inhibitory Effect of Curcumin Nanoparticles and Free Curcumin in Human Telomerase Reverse Transcriptase Gene Expression in Breast Cancer. Adv. Pharm. Bull. 2013, 3, 127–130. [Google Scholar]
  309. Pourhassan, M.; Zarghami, N.; Rahmati, M.; Alibakhshi, A.; Ranjbari, J.; Mohammad, P.; Nosratollah, Z.; Abbas, A.; Javad, R. The inhibitory effect of Curcuma longa extract on telomerase activity in A549 lung cancer cell line. Afr. J. Biotechnol. 2010, 9, 912–919. [Google Scholar] [CrossRef]
  310. Tavakoli, F.; Jahanban-Esfahlan, R.; Seidi, K.; Jabbari, M.; Behzadi, R.; Soltanahmadid, Y.P.; Zarghami, N. Effects of nano-encapsulated curcumin-chrysin on telomerase, MMPs and TIMPs gene expression in mouse B16F10 melanoma tumour model. Artif. Cells Nanomed. Biotechnol. 2018, 46, 75–86. [Google Scholar] [CrossRef] [Green Version]
  311. Khokha, R. Suppression of the Tumorigenic and Metastatic Abilities of Murine B16-F10 Melanoma Cells In Vivo by the Overexpression of the Tissue Inhibitor of the Metalloproteinases-1. J. Natl. Cancer Inst. 1994, 86, 299–304. [Google Scholar] [CrossRef]
  312. Komath, S.; Garg, A.; Wahajuddin, M. Development and evaluation of Chrysin-Phospholipid complex loaded solid lipid nanoparticles-storage stability and in vitro anti-cancer activity. J. Microencapsul. 2018, 35, 600–617. [Google Scholar] [CrossRef] [PubMed]
  313. Mohammadi, Z.; Zak, M.S.; Seidi, K.; Barati, M.; Akbarzadeh, A.; Zarghami, N. The Effect of Chrysin Loaded PLGA-PEG on Metalloproteinase Gene Expression in Mouse 4T1 Tumor Model. Drug Res. 2017, 67, 211–216. [Google Scholar] [CrossRef] [PubMed]
  314. Khaledi, S.; Jafari, S.; Hamidi, S.; Molavi, O.; Davaran, S. Preparation and characterization of PLGA-PEG-PLGA polymeric nanoparticles for co-delivery of 5-Fluorouracil and Chrysin. J. Biomater. Sci. Polym. Ed. 2020, 31, 1107–1126. [Google Scholar] [CrossRef] [PubMed]
  315. Ferrado, J.B.; Perez, A.A.; Ruiz, M.C.; León, I.E.; Santiago, L.G.; Rubin, A.A.P. Chrysin-loaded bovine serum albumin particles as bioactive nanosupplements. Food Funct. 2020, 11, 6007–6019. [Google Scholar] [CrossRef]
  316. Lotfi-Attari, J.; Soltanahmadid, Y.P.-; Dadashpour, M.; Alipour, S.; Farajzadeh, R.; Javidfar, S.; Zarghami, N. Co-Delivery of Curcumin and Chrysin by Polymeric Nanoparticles Inhibit Synergistically Growth and hTERT Gene Expression in Human Colorectal Cancer Cells. Nutr. Cancer 2017, 69, 1290–1299. [Google Scholar] [CrossRef]
  317. Mohammadinejad, S.; Akbarzadeh, A.; Rahmati-Yamchi, M.; Hatam, S.; Kachalaki, S.; Zohreh, S.; Zarghami, N. Preparation and Evaluation of Chrysin Encapsulated in PLGA-PEG Nanoparticles in the T47-D Breast Cancer Cell Line. Asian Pac. J. Cancer Prev. 2015, 16, 3753–3758. [Google Scholar] [CrossRef] [Green Version]
  318. Anari, E.; Akbarzadeh, A.; Zarghami, N. Chrysin-loaded PLGA-PEG nanoparticles designed for enhanced effect on the breast cancer cell line. Artif. Cells Nanomed. Biotechnol. 2015, 44, 1410–1416. [Google Scholar] [CrossRef]
  319. Javan, N.; Ansari, M.H.K.; Dadashpour, M.; Khojastehfard, M.; Bastami, M.; Rahmati-Yamchi, M.; Zarghami, N. Synergistic Antiproliferative Effects of Co-nanoencapsulated Curcumin and Chrysin on MDA-MB-231 Breast Cancer Cells Through Upregulating miR-132 and miR-502c. Nutr. Cancer 2019, 71, 1201–1213. [Google Scholar] [CrossRef]
  320. Kim, K.M.; Lim, H.K.; Shim, S.H.; Jung, J. Improved chemotherapeutic efficacy of injectable chrysin encapsulated by copolymer nanoparticles. Int. J. Nanomed. 2017, 12, 1917–1925. [Google Scholar] [CrossRef] [Green Version]
  321. Baidya, D.; Kushwaha, J.; Mahadik, K.; Patil, S. Chrysin-loaded folate conjugated PF127-F68 mixed micelles with enhanced oral bioavailability and anticancer activity against human breast cancer cells. Drug Dev. Ind. Pharm. 2019, 45, 852–860. [Google Scholar] [CrossRef]
  322. Mohammadian, F.; Soltanahmadid, Y.P.; Mofarrah, M.; Dastani-Habashi, M.; Zarghami, N. Down regulation of miR-18a, miR-21 and miR-221 genes in gastric cancer cell line by chrysin-loaded PLGA-PEG nanoparticles. Artif. Cells Nanomed. Biotechnol. 2016, 44, 1972–1978. [Google Scholar] [CrossRef] [PubMed]
  323. Yadav, S.; Singh, J.D. Synthesis and preliminary biological evaluation for the anticancer activity of organochalcogen (S/se) tethered chrysin-based organometallic RuII(η6-p-cymene) complexes. J. Biomol. Struct. Dyn. 2018, 37, 3337–3353. [Google Scholar] [CrossRef] [PubMed]
  324. Wang, H.-J.; Zhou, Y.-Y.; Liu, X.-L.; Zhang, W.-H.; Chen, S.; Zhou, Y.; Liu, X.-W. Regioselective synthesis and evaluation of 2-amino 3-cyano chromene-chrysin hybrids as potential anticancer agents. Bioorg. Med. Chem. Lett. 2020, 30, 127087. [Google Scholar] [CrossRef] [PubMed]
  325. Zhang, W.-H.; Chen, S.; Liu, X.-L.; Lin, B.; Zhou, Y. Study on antitumor activities of the chrysin-chromene-spirooxindole on Lewis lung carcinoma C57BL/6 mice in vivo. Bioorg. Med. Chem. Lett. 2020, 30, 127410. [Google Scholar] [CrossRef] [PubMed]
  326. Al-Oudat, B.A.; Ramapuram, H.; Malla, S.; Audat, S.A.; Hussein, N.; Len, J.M.; Kumari, S.; Bedi, M.F.; Ashby, J.C.R.; Tiwari, A.K. Novel Chrysin-De-Allyl PAC-1 Hybrid Analogues as Anticancer Compounds: Design, Synthesis, and Biological Evaluation. Molecules 2020, 25, 3063. [Google Scholar] [CrossRef]
  327. Mayer, S.; Keglevich, P.; Ábrányi-Balogh, P.; Szigetvári, Á.; Dékány, M.; Szántay, J.C.; Hazai, L. Synthesis and In Vitro Anticancer Evaluation of Novel Chrysin and 7-Aminochrysin Derivatives. Molecules 2020, 25, 888. [Google Scholar] [CrossRef] [Green Version]
  328. Xuan, H.; Zhang, J.-H.; Wang, Y.-H.; Fu, C.-L.; Zhang, W. Anti-tumor activity evaluation of novel chrysin–organotin compound in MCF-7 cells. Bioorg. Med. Chem. Lett. 2016, 26, 570–574. [Google Scholar] [CrossRef]
  329. Halevas, E.; Mavroidi, B.; Antonoglou, O.; Hatzidimitriou, A.; Sagnou, M.; Pantazaki, A.A.; Litsardakis, G.; Pelecanou, M. Structurally characterized gallium–chrysin complexes with anticancer potential. Dalton Trans. 2020, 49, 2734–2746. [Google Scholar] [CrossRef]
  330. Zhang, H.M.; Li, H.; Wang, G.X.; Wang, J.; Xiang, Y.; Huang, Y.; Shen, C.; Dai, Z.T.; Li, J.P.; Zhang, T.C.; et al. MKL1/miR-5100/CAAP1 loop regulates autophagy and apoptosis in gastric cancer cells. Neoplasia 2020, 22, 220–230. [Google Scholar] [CrossRef]
  331. Zhang, L.; Liu, X.; Song, L.; Zhai, H.; Chang, C. MAP7 promotes migration and invasion and progression of human cervical cancer through modulating the autophagy. Cancer Cell Int. 2020, 20, 17–18. [Google Scholar] [CrossRef]
  332. Liang, G.; Ling, Y.; Mehrpour, M.; Saw, P.E.; Liu, Z.; Tan, W.; Tian, Z.; Zhong, W.; Lin, W.; Luo, Q.; et al. Autophagy-associated circRNA circCDYL augments autophagy and promotes breast cancer progression. Mol. Cancer 2020, 19, 65. [Google Scholar] [CrossRef] [PubMed]
  333. Pan, Z.; Wu, C.; Li, Y.; Li, H.; An, Y.; Wang, G.; Dai, J.; Wang, Q. LncRNA DANCR silence inhibits SOX5-medicated progression and autophagy in osteosarcoma via regulating miR-216a-5p. Biomed. Pharmacother. 2019, 122, 109707. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structure of chrysin.
Figure 1. Chemical structure of chrysin.
Biomolecules 10 01374 g001
Figure 2. A schematic representation of the health-promoting effects of chrysin in pre-clinical experiments.
Figure 2. A schematic representation of the health-promoting effects of chrysin in pre-clinical experiments.
Biomolecules 10 01374 g002
Figure 3. Mechanisms involved in the anti-tumor activity of chrysin against different cancers. IL-1β, interleukin-1β; IL-6, interleukn-6; TNF-α, tumor necrosis factor-α; NF-κB, nuclear factor-kappaB; ROS, reactive oxygen species; ER, endoplasmic reticulum; Mcl-1, myeloid cell leukemia-1; EGFR, epidermal growth factor receptor; VEGF, vascular endothelial growth factor; miR, microRNA; MMP, matrix metalloproteinase; TLR4, toll-like receptor 4; mTOR, mammalian target of rapamycin.
Figure 3. Mechanisms involved in the anti-tumor activity of chrysin against different cancers. IL-1β, interleukin-1β; IL-6, interleukn-6; TNF-α, tumor necrosis factor-α; NF-κB, nuclear factor-kappaB; ROS, reactive oxygen species; ER, endoplasmic reticulum; Mcl-1, myeloid cell leukemia-1; EGFR, epidermal growth factor receptor; VEGF, vascular endothelial growth factor; miR, microRNA; MMP, matrix metalloproteinase; TLR4, toll-like receptor 4; mTOR, mammalian target of rapamycin.
Biomolecules 10 01374 g003
Figure 4. Chrysin-loaded nanoparticles in cancer therapy.
Figure 4. Chrysin-loaded nanoparticles in cancer therapy.
Biomolecules 10 01374 g004
Table 1. Various pharmacological activities of chrysin in treating diseases.
Table 1. Various pharmacological activities of chrysin in treating diseases.
Therapeutic Effect/DiseaseIn Vitro/In VivoCell Line/Animal ModelDose (In Vivo)/Concentration (In Vitro)Duration of ExperimentAdministration RouteOutcomesRefs
Anti-hypertensionIn vivoRat100 mg/kg18 weeksOral administrationDecreasing systolic and diastolic pressures
Reducing insulin, angiotensin II and tiacylglycerols levels
[85]
NeuroprotectiveIn vivoRat10 and 30 mg/kg8 weeksOral gavageImproving memory impairment
Enhancing neuronal cell survival
Reducing hippocampal neurogenesis depletion
[86]
NeuroprotectiveIn vivoRat10, 30, and 100 mg/kg3 weeksOral administrationEnhancing GPx activity and number of surviving cells in hippocampus
Reducing MDA, NO and PGE2 levels
Improving passive avoidance memory
[87]
CardioprotectiveIn vitroCardiomyocyte10, 50, and 100 µM3 h-Decreasing aluminium-phosphide-mediated oxidative stress
Reducing mitochondrial damage
Improving mitochondrial function
[88]
Renoprotective
Hepatoprotective
In vivoRat100 mg/kg--Reinforcing antioxidant defense system via up-regulating GSH and SOD activities
Reducing lipid peroxidation
Decreasing inflammation via TNF-α down-regulation
[89]
Renoprotective
Hepatoprotective
In vivoRat25 and 50 mg/kg7 daysOral administrationReducing AST, ALT, ALP, urea, creatinine, MDA and hepatorenal deterioration
Enhancing SOD, CAT, and GPx activities
Apoptosis inhibition via Bcl-2 up-regulation and Bax down-regulation
Reducing inflammation via NF-κB down-regulation
[90]
Anti-diabeticIn vitroChorioretinal endothelial cells1, 3, 10, 30, and 50 µM24 h-Reducing Akt, ERK, MMP-2, and VEGF expressions[91]
Anti-diabeticIn vivoRat model of type I diabetes50 and 100 mg/kg28 daysOral gavageReducing oxidative stress index
Enhancing glutathione levels
[92]
Gastric healingIn vivoMouse model of gastric ulcer via ethanol10, 50, and 100 mg/kg7 and 14 daysOral administrationApoptosis inhibition via caspase-3 down-regulation
Reducing macroscopic lesions
Enhancing catalase activity
Improving inflammation via COX-2 down-regulation
[93]
Table 2. Chrysin administration in treating various cancers.
Table 2. Chrysin administration in treating various cancers.
Cancer TypeIn Vitro/In VivoCell Line/Animal ModelDose (In Vivo)/Concentration (In Vitro)Period of ExperimentAdministration RouteOutcomesRefs
Prostate cancerIn vitroDU145 and PC-3 cell lines12.5, 25 and 50 µM--Induction of mitochondrion- and ER-mediated apoptosis
Cell cycle arrest
Down-regulation of MAPK and PI3K/Akt signaling pathways
Impairing proliferation of PC cells
[144]
Gastric cancerIn vitroMKN45 cells
Mouse model of GC (created by CRIPSR/Cas9)
10, 20, 40, 80 and 160 µM
20 mg/kg
12, 24 and 45 h
14 days
Oral gavageSuppressing migration
Apoptosis induction
Enhancing TET1 expression
[165]
Lung cancerIn vitroA549 cells2 and 5 µM4 h-Down-regulation of MyD88 and TLR4
Inhibition of inflammation via NF-κB down-regulation
Suppressing survival and metastasis
[134]
Cervical cancerIn vitroHeLa cells5, 10, 20 and 40 µM0.5, 3, 6, 12 and 24 h-Down-regulation of NF-κB signaling pathway
Inhibition of Twist/EMT axis
Suppressing metastasis of cervical cancer
[182]
Breast cancerIn vitroT47D breast cancer cells20, 40, 60, 80, 100 and 120 µM48 h-Disrupting proliferation of cancer cells via down-regulation of cyclin D1 and hTERT[105]
Hepatocellular carcinomaIn vitro
In vivo
Normal human hepatic cell LO2 and HepG2, Hep3B, Huh-7, HCC-LM3, Bel-7402 and SMMC-7721
Tumor xenografts
15, 30, and 60 µM
30 mg/kg
24, 48 and 72 hIntraperitoneal injectionDown-regulation of HK-2
Suppressing glycolysis
Apoptosis induction
[194]
Breast cancer
Cervical cancer
In vitroHeLa cells
MCF-7 cells
15, 20, 25 and 30 µM30 min-Significant reduction in survival of cancer cells
Inducing both intrinsic and extrinsic apoptotic pathways
P53-dependent apoptosis
[246]
Ovarian cancerIn vitroSKOV3 cell line5, 10 and 20 µmol/L--Decreasing the viability of cancer cells in a dose-dependent manner
Down-regulation of CK2α, CD133 and CD44
Suppressing sphere formation capability
[247]
Breast cancerIn vitroMDA-MB-23110 µM24 and 48 h-Inhibition of EGFR
Reducing migration, growth and sphere formation ability of cancer cells
[109]
Breast cancerIn vitro
In vivo
4T1 mouse breast cancer cells
Balb/c mice implanted with 4T1 cells
60–100 µM
250 mg/kg
30 min
18 days
Oral administrationSuppressing lung metastasis
Down-regulation of VEGF, and STAT3
Inhibiting proliferation
[119]
Prostate cancerIn vitroHuman prostate cancer cell line PC-310, 20, 30, and 40 µM24, 48 and 72 h-Reducing the viability of cancer cells in a time- and dose-dependent manner
Apoptosis induction
[248]
Cervical cancerIn vitroHuman cervical epidermoid carcinoma cell line ME180, and human cervical carcinoma cell lines HeLa, BU25TK− and SiHa0–160 mg/mL--Apoptosis induction via caspase-3, caspase-9, and Bax up-regulation
Stimulating mitochondrial dysfunction
Cell cycle arrest induction
[184]
Liver cancerIn vitroHepatocellular carcinoma cells5–100 µM15, 30, 45 and 60 min-Mitochondrial dysfunction
Cytochrome c release into the cytoplasm
Apoptosis induction
[195]
Breast cancerIn vitroMDA-MB-231 and MCF-7 cells3–12 µM--Reducing the viability of cancer cells
Apoptosis induction via capase-3 and caspase-7 up-regulation
[249]
MelanomaIn vitro
In vivo
B16F10 cells
Melanoma-bearing mice
12.5, 25, 50, and 100 µM
50 mg/kg
24 and 48 h
21 days
-Induction of cell cycle arrest at G2/m phase
Reducing tumor growth in vivo
Promoting the anti-tumor activity of immune cells, such as macrophages and natural killer cells
[211]
Oral squamous cell carcinomaIn vitroOral squamous carcinoma KB cell line1, 2, 4, 8, 16, and 32 µmol/L24 h-Suppressing proliferation in a dose-dependent manner
Apoptosis induction via capase-3 and caspase-7 up-regulation
Inducing mitochondrial dysfunction
Reducing the viability via down-regulation of PI3K/Akt signaling pathways
[250]
Bladder cancerIn vitroHuman bladder cancer cell lines T-24 and 5637 and the non-malignant immortalized urothelial SV-HUC-1 cells 20, 40 and 80 µM24 h-Induction of ER stress via UPR activation
Stimulating intrinsic pathway of apoptosis via caspase-3 and caspase-9 up-regulation
Inhibition of STAT3 signaling pathway
[251]
MelanomaIn vitroHuman melanoma A375.S2 cell line5, 10 and 15 µM24 and 48 h-Impairing metastasis via VEGF, MMP-2, and N-cadherin down-regulation
Enhancing E-cadherin expression
Down-regulation of PI3K/Akt and NF-κB pathways in suppressing cancer proliferation
[182]
Colorectal cancerIn vitroSW48, SW480, and SW620 CRC cells 5–50 µM24 h-Enhancing ROS generation
mTOR down-regulation
Elevating LC-3II levels
Autophagy induction
Impairing cancer cell viability
[227]
Breast cancerIn vitroMCF-7 cells20 and 30 µM48 and 72 h-Anti-proliferative activity in a dose- and time-dependent manner
Apoptosis induction
[252]
Cervical cancerIn vitroHeLa cells0–10 µM12–48 h-Stimulating apoptosis and cell cycle arrest
Down-regulation of COX-2 expression
[253]
Colon cancerIn vitroHT-29 cells12.5, 25, 50, and 100 µg/mL--Induction of apoptosis via mitochondrial dysfunction
Irradiation combined with chrysin exerts a synergistic effect
[235]
Thyroid carcinomaIn vitro
In vivo
HTh7 and KAT18 cells25, 50, and 75 µM
75 mg/kg
2–6 days
21 days
Oral gavageReducing the viability and growth via up-regulation of Notch1 and its down-stream target, Hes1[254]
Hepatocellular carcinomaIn vitroSMMC-7721 cells10, 20 and 40 µM24 and 48 h-Reducing sphere formation via STAT3 down-regulation[203]
Breast cancerIn vitroMCF-7 cells40 µM8 h-Decreasing cell viability by p53 activation through ATM-ChK2 axis
Lack of DNA damage
[255]
Tongue squamous cell carcinomaIn vitroCAL-27 cells5, 25, 55 and 80 µM24 h-Apoptosis induction via caspase-3 and caspase-9 up-regulation[256]
Choriocarcinoma cellsIn vitroJAR and JEG3 cells0–100 µM24 h-Suppressing cell viability in a dose-dependent manner
Inducing cell death via promoting ROS production and changing mitochondrial membrane potential
[257]
Colorectal cancerIn vitroHCT116 cells20, 30, 40 and 50 µM36 h-Cell cycle arrest
Migration inhibition
PARPα up-regulation
CYP2S1 and CYP1B1 induction
[245]
Colon cancerIn vitro
In vivo
CT26 cells
Allograft colon carcinoma model
10–200 µg/mL
0–10 mg/kg
24 and 48 h
28 days
Oral administrationReducing tumor growth
Induction of apoptosis via caspase-3 and caspase-9 up-regulation
[258]
Table 3. Chrysin-loaded nanoparticles in cancer therapy.
Table 3. Chrysin-loaded nanoparticles in cancer therapy.
NanovehicleCancer TypeIn Vitro/In VivoCell Line/Animal ModelParticle Size (nm)Zeta Potential (mV)Encapsulation Efficiency (%)OutcomesRefs
MicelleColorectal cancerIn vitroHuman-derived epithelial colorectal cancer cell lines HT-2972–142+10.177 (Docetaxel)
44 (chrysin)
Enhanced cellular uptake
Effective inhibition of cancer stem cell migration
[292]
Polymeric micellesBreast cancerIn vitroMCF-7 cells55−2.787.6 (methotrexate)
86.5 (chrysin)
Enhancing efficacy of chrysin and methotrexate in breast cancer therapy via promoting cellular uptake[293]
DendrimerOvarian cancerIn vitroSerous carcinoma (OSC) cell lines (OVCAR3 HTB-161™ and OVCAR8 CVCL_1629™) and a clear cell carcinoma (OCCC) cell line (ES2 CRL-1978™)---Selective targeting of cancer cells by folate functionalization of dendrimers
High cellular uptake
Remarkable decrease in survival of cancer cells
[297]
Polymeric nanoparticlesBreast cancerIn vitroT47D breast cancer cell line75-99.89Higher cytotoxicity against breast cancer cells compared to chrysin alone[304]
PLGA-PEG nanoparticlesBreast cancerIn vitroT47-D breast cancer cell line20–75-70High cytotoxicity
Excellent cellular uptake and encapsulation efficiency
[317]
PLGA-PEG nanoparticlesColorectal cancerIn vitroSW480 cells50–140 nm- Higher cytotoxicity compared to chrysin and curcumin alone
hTERT down-regulation
[307]
PLGA-PEG nanoparticlesMelanomaIn vivoC57B16 mice bearing B16F10 melanoma tumours285−3.778.27 (curcumin)
83.5 (chrysin)
Enhancing expression of TIMP-1 and TIMP-2
Down-regulation of MMP-2 and MMP-9
Suppressing metastasis of cancer cells
[310]
Solid lipid nanoparticlesBreast cancerIn vitroMCF-7 cellsBelow 500−20 to −47More than 90%High stability and promoting the anti-tumor activity of chrysin[312]
PLGA-PEG nanoparticlesBreast cancerIn vitroT47D cells70–300-99.89Accumulation in breast cancer cells
High cytotoxicity
[318]
PLGA-PEG nanoparticlesBreast cancerIn vitroMDA-MB-231 cells305−3.880.22 (curcumin)
85.25 (chrysin)
Synergistic effect
Cell cycle arrest at G2/M phase
Apoptosis induction
Up-regulation of miR-132 and miR-502c
[319]
Copolymer nanoparticleLung cancerIn vitro
In vivo
A549 cells
Mice bearing an A549-derived tumor
77−2.2246.96Enhanced cytotoxicity
More potential in exerting tumor growth delay
[320]
MicelleBreast cancerIn vitroMCF-7 cells152–420−21.652–89Promoting bioavailability of chrysin
Exerting a 5-fold increase in anti-tumor activity
[321]
PLGA-PEG nanoparticlesGastric cancerIn vitroAGS cells70–300-98.6Decreasing cell survival via down-regulation of miR-18a, miR-21, and miR-221[322]

Share and Cite

MDPI and ACS Style

Moghadam, E.R.; Ang, H.L.; Asnaf, S.E.; Zabolian, A.; Saleki, H.; Yavari, M.; Esmaeili, H.; Zarrabi, A.; Ashrafizadeh, M.; Kumar, A.P. Broad-Spectrum Preclinical Antitumor Activity of Chrysin: Current Trends and Future Perspectives. Biomolecules 2020, 10, 1374. https://0-doi-org.brum.beds.ac.uk/10.3390/biom10101374

AMA Style

Moghadam ER, Ang HL, Asnaf SE, Zabolian A, Saleki H, Yavari M, Esmaeili H, Zarrabi A, Ashrafizadeh M, Kumar AP. Broad-Spectrum Preclinical Antitumor Activity of Chrysin: Current Trends and Future Perspectives. Biomolecules. 2020; 10(10):1374. https://0-doi-org.brum.beds.ac.uk/10.3390/biom10101374

Chicago/Turabian Style

Moghadam, Ebrahim Rahmani, Hui Li Ang, Sholeh Etehad Asnaf, Amirhossein Zabolian, Hossein Saleki, Mohammad Yavari, Hossein Esmaeili, Ali Zarrabi, Milad Ashrafizadeh, and Alan Prem Kumar. 2020. "Broad-Spectrum Preclinical Antitumor Activity of Chrysin: Current Trends and Future Perspectives" Biomolecules 10, no. 10: 1374. https://0-doi-org.brum.beds.ac.uk/10.3390/biom10101374

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop