Next Article in Journal
Microbial Natural Products with Antiviral Activities, Including Anti-SARS-CoV-2: A Review
Previous Article in Journal
A Strategy for Identification and Structural Characterization of Compounds from Plantago asiatica L. by Liquid Chromatography-Mass Spectrometry Combined with Ion Mobility Spectrometry
Previous Article in Special Issue
Exploring the Phytochemicals and Anti-Cancer Potential of the Members of Fabaceae Family: A Comprehensive Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pharmacological Properties of 4′, 5, 7-Trihydroxyflavone (Apigenin) and Its Impact on Cell Signaling Pathways

by
Rameesha Abid
1,2,*,
Shakira Ghazanfar
2,
Arshad Farid
3,*,
Samra Muhammad Sulaman
4,
Maryam Idrees
2,5,
Radwa Abdallnasser Amen
6,
Muhammad Muzammal
3,
Muhammad Khurram Shahzad
7,
Mohamed Omar Mohamed
8,
Alaa Ashraf Khaled
9,
Waqas Safir
10,
Ifra Ghori
11,
Abdelbaset Mohamed Elasbali
12,* and
Bandar Alharbi
13
1
Department of Biotechnology, University of Sialkot, Sialkot 51310, Pakistan
2
National Institute for Genomics and Advanced Biotechnology (NIGAB), National Agricultural Research Center, Islamabad 44100, Pakistan
3
Gomal Center of Biochemistry and Biotechnology, Gomal University, Dera Ismail Khan 29050, Pakistan
4
Biotechnology Department, University of Gujrat, Punjab 50700, Pakistan
5
Department of Microbiology, Quaid-i-Azam University, Islamabad 45320, Pakistan
6
Biotechnology Department, Faculty of Science, Cairo University, Cairo 424010, Egypt
7
Biotechnology and Bioinformatics Department, International Islamic University, Islamabad 44100, Pakistan
8
Faculty of Agriculture, Ain Shams University, Cairo 424010, Egypt
9
Faculty of Agriculture, Benha University, Benha 13511, Egypt
10
College of Life Science and Technology, Xinjiang University, Urumqi 830046, China
11
Department of Biotechnology, Fatima Jinnah Women University, Rawalpindi 46000, Pakistan
12
Department of Clinical Laboratory Science, College of Applied Sciences-Qurayyat, Jouf University, Sakaka 72388, Saudi Arabia
13
Department of Medical Laboratory, College of Applied Medical Science, University of Hail, Hail 81481, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(13), 4304; https://doi.org/10.3390/molecules27134304
Submission received: 18 May 2022 / Revised: 23 June 2022 / Accepted: 1 July 2022 / Published: 4 July 2022

Abstract

:
Plant bioactive compounds, particularly apigenin, have therapeutic potential and functional activities that aid in the prevention of infectious diseases in many mammalian bodies and promote tumor growth inhibition. Apigenin is a flavonoid with low toxicities and numerous bioactive properties due to which it has been considered as a traditional medicine for decades. Apigenin shows synergistic effects in combined treatment with sorafenib in the HepG2 human cell line (HCC) in less time and statistically reduces the viability of tumor cells, migration, gene expression and apoptosis. The combination of anti-cancerous drugs with apigenin has shown health promoting potential against various cancers. It can prevent cell mobility, maintain the cell cycle and stimulate the immune system. Apigenin also suppresses mTOR activity and raises the UVB-induced phagocytosis and reduces the cancerous cell proliferation and growth. It also has a high safety threshold, and active (anti-cancer) doses can be gained by consuming a vegetable and apigenin rich diet. Apigenin also boosted autophagosome formation, decreased cell proliferation and activated autophagy by preventing the activity of the PI3K pathway, specifically in HepG2 cells. This paper provides an updated overview of apigenin’s beneficial anti-inflammatory, antibacterial, antiviral, and anticancer effects, making it a step in the right direction for therapeutics. This study also critically analyzed the effect of apigenin on cancer cell signaling pathways including the PI3K/AKT/MTOR, JAK/STAT, NF-κB and ERK/MAPK pathways.

1. Introduction

Different bioactive compounds of plants have manifested many functional activities that could be an astonishing factor in the prevention of various diseases. Flavonoids are a large group of natural polyphenols which include isoflavonoids, anthocyanidins, flavonols, flavanones and flavonols. They are a class of polyphenolic compounds and are distinguished by broad biological activities that occur in many mammalian bodies, including inhibition of tumor growth [1]. Among all of the polyphenols, the five most common plant polyphenols are kaempferol, quercetin, myricetin, kaempferol, apigenin, and luteolin, which are abundantly found in over 6000 different plants located in Western Asia, Europe, Australia and Britain [2]. The molecular formula of apigenin is C15H10O5. It is a trihydroxyflavone that is substituted by OH functional group at the 4’, 5 and 7 position. It has reduced water solubility and greater permeability due to which it can easily pass through the plasma membrane of the host body. Moreover, it is lipophilic in nature and can survive in the gastrointestinal tract of the host under low acidic environment [3]. Apigenin is one of the most common monomeric flavonoids found in chamomile (Matricaria recutita), some green vegetables, artichokes, celery, parsley, herbs, sorghum and dried oregano. Chamomile is traditionally used to treat insomnia and anxiety. It shows a wide range of beneficial bioactivities in combating different diseases including cancer [4,5]. It also helps to induce autophagy in leukemia cells and acts as an antineoplastic agent.
A medical study conducted simultaneously among healthy and ovarian cancer affected women revealed that flavonoid intake has no evident connection with the cause of ovarian cancer. However, a significant decrease in the risk and potential of ovarian cancer was observed due to regular apigenin intake. After a course of medical experiments, it has been inferred that apigenin together with luteolin works as a potent chemotherapeutic agent [6].
Recently, the scientific community has explored the most promising health benefits of apigenin [7]. Apigenin has the potential to cause less intrinsic cytotoxicity and showed arrestive effects on cancerous cells as compared to other conventional flavonoids [8]. Studies have suggested that apigenin has a strong background in the prevention of numerous infections by using different experimental analyses [5]. The anti-inflammatory and antioxidant potential of apigenin [9,10] induces apoptosis and proliferation at a cellular level in host organisms and halts the overexpression of cancer-causing genes [11,12]. Apigenin regulates cell death by activating various intrinsic apoptotic pathways and caspase-3 activity, resulting in forming apoptotic protease activating factor 1(APAF) by liberating one of the components of the electron transport chain named cytochrome C [13]. This flavonoid is thought to induce antioxidant potential by activating different signaling cascades and inhibiting the activity of the NF-κB pathway. To enhance the antioxidant potential, apigenin activates antioxidant enzymes including erythrocyte superoxide dismutase, catalase, phase II detoxification enzymes and Glutathione Peroxidase (GSH-synthase) [14,15]. Apigenin not only reduced the expression of different tumor-causing genes such as TNF-α, IL-6 and CD40, but also increased the activity of the tumor-suppressing STAT1 gene via IFN-γ gene inhibition [16].
Research on mice demonstrated that apigenin reduces cholesterol levels, fatty acid production levels and obesity in mice. It regulates nicotinamide adenine dinucleotide levels and stabilizes glucose levels in the human body [17,18]. It decreases inflammation and pain by inhibiting different cellular processes and pro-inflammatory pathways. It decreases the activity and signaling of NF-κB [19], JAK2 and STAT3 [20] pathways and suppresses the expression of different cytokines [21] including Th2 cytokines, IL-4, IL-10 [22], interleukin-1β and NLRP3 genes [23]. Apigenin shows hypoglycemic effects. Its bioactive compounds can reduce insulin levels and helps to maintain the normal glucose levels in the body. Apigenin also has the potential to enhance wound and bone healing effects [24]. It is an important nutraceutical plant compound that has a noticeable effect on different chronic diseases such as Alzheimer’s, stroke, diabetes, cancer, anxiety and depression; it also has anti-microbial, anti-tumorigenic, anti-mutagenic and antiviral properties [25,26]. Recently, the scientific community has noticed the growing challenge of antimicrobial resistance (AMR) to the world. The number of multidrug-resistant organisms are increasing due to the inappropriate use of antibiotics. Although international healthcare authorities are trying to create awareness campaigns through antibiotic stewardship programs as, the risk of AMR and associated comorbidities are constantly developing. To overcome this problem, medicinal plants, especially apigenin and its bioactive compounds, can be regarded as an alternative way to treat different metabolic diseases. Literature describing the medicinal potential of apigenin has been published previously. However, an extensive review is needed to evaluate the medicinal importance of apigenin through different meta-analyses. The main objective of this review paper is to highlight the biological and pharmacological properties of apigenin which can be used to completely eradicate and obstruct the pathogenicity of various diseases. In this review, we also investigated the most promising health benefits of apigenin and its role in the cancer signaling pathways.

2. Anti-Inflammatory and Antioxidant Potential of Apigenin

Numerous studies have shown that apigenin has gained the spotlight of the scientific community because of its anti-inflammatory and antioxidant properties. The antioxidant properties of apigenin are mediated via antioxidant enzymes including superoxide dismutases (SOD), glutathione reductase (GR) and catalase (CAT) (Figure 1) [27]. Parsley having 3.73–4.49 mg of apigenin, consumed by humans for fifteen days, resulted in elevated levels of glutathione reductase and superoxide dismutases as compared to a less flavonoid-based human diet. Another study proved that adequate apigenin doses (10, 20, and 40 mg/kg) give safety to rat livers, decrease lipid peroxidation (LPO) and secrete blood serum enzyme markers via lactate dehydrogenase, alkaline phosphatase, alanine aminotransferase, aspartate transaminase and protect rats from oxidative and membrane protein damage [28,29]. This flavonoid can suppress the activity of different diseases including cancer, those of the cardiovascular system, and diabetes by acting as a therapeutic agent, and helps to promote host health [30,31]. Apigenin has the potency to produce an enzyme-inhibitor complex and block the inflammatory activities of different enzymes [32]. A study on mouse white blood cells demonstrated that apigenin inhibits the lipopolysaccharide-induced COX-2 and NOS inhibitors activity, resulting in an anti-inflammatory response in their bodies [32] (Figure 1). Multiple cytokines, monocyte inflammatory protein (MIP-1α), intracellular cell adhesion molecules (ICAMS), and monocyte chemotactic protein (MCP-1α) also cause inflammation in the host body. Apigenin is responsible for down-regulating the expression of these cytokines via extracellular signal-regulated kinases (ERK) and mitogen-activated protein kinase (MAPK) (Figure 1) [33]. Reactive oxygen species (ROS) like alkoxyl, hydroxyl, peroxyl, nitric acid and superoxide can attack the biological molecules, resulting in oxidative damage in the human body [34,35]. These ROS help to activate multiple transcription factors like PPAR-γ, HIF-1α, Sp-1, STAT-3, AP-1 and Nrf2, and enhance different genes including inflammatory cytokines and various chronic diseases mediated through increases in their expression levels [36,37] (Figure 1). Apigenin can also treat these chronic diseases by suppressing the expression of different genes [29]. In human peripheral blood lymphocytes (HPBL), apigenin not only reduces the probability of apoptosis and mitochondrial membrane potential depolarization (ΔΨm), but also stabilizes the antioxidant status of the cell and protects the cell from oxidative damage and lipid peroxidation [38] (Figure 1).

3. Antibacterial Potential of Apigenin

Generally, phytochemicals like apigenin have noticeable activity against microbial and bacterial infections [26]. Many studies have investigated the potential role of apigenin against microbes. It is effective in treating amnesia and various other disorders and has antioxidant, anticancer and anti-inflammatory characteristics.
Flavones fight against bacterial infections through a fascinating mechanism. They inhibits the nucleic acid synthetization and the functionality of the cytoplasmic membrane by forming beta-barrel proteins and biofilms in bacterial species. They also interact with pivotal enzymes. When it comes to apigenin it inhibits the DNA (deoxyribonucleic acid) and biofilm formation of gram-negative Escherichia coli bacteria (Figure 1). Studies have shown that the antibacterial potential of flavones and phytochemicals is dependent on the selected strains of bacteria. It is worth noting that the relationship between the structural and antimicrobial activity of flavones is a subject of focus for future research [28].
Some studies have mentioned that apigenin is strongly recommended for the treatment of oral bacteria, as it possesses significant antibacterial properties. The effectiveness of apigenin with antibiotics against these bacteria increased gradually as compared to the antibiotics treatment alone, so there is a synergic effect between apigenin and antibiotics [29]. Most of these studies and experiments have been done on strain-specific bacteria in aerobic conditions, and it has been observed that apigenin can inhibit bacterial colonies and decrease their toxicity [30]. A study claimed that apigenin is abundantly present in the leaves of Portulaca oleracea. This plant is known because of its antioxidant, antitoxic, antimicrobial and hypoglycemic activities, which produce beneficial effects on the host body.
Apigenin can be administered in the form of tablets, capsules and oral suspensions [31,32,33,34]. Research has suggested that the most significant antibacterial activity of apigenin has been seen against gram-negative Proteus mirabilis [35]. Apigenin has the potential to inhibit cell proliferation via the P13k/AKT cell signaling pathway [36,37]. It shows antibacterial activities against various gram-negative and positive bacterial species [32,33,34,38].

4. Antiviral Beneficiary Effects

As a result of the ubiquitous spread of viral infections with no effective treatment and the continuous arrival of resistant viral strains, the improvement in the health care systems and to treat infectious diseases with novel agents is the exigency of the time. It is necessary to standardize effective and functional medications to control future outbreaks [39]. Flavonoids are effective against various DNA and RNA enveloped and non-enveloped viruses. They can prevent viruses from attaching to cells and infecting other cells by interacting with different viral replication stages and also inhibit the process of protein synthesis. Flavonoids have been discovered to inhibit viruses in a variety of ways [40].
Apigenin is a non-virulent and non-hazardous active flavonoid [41]. Studies have shown that the reduced expression of mature miR122 by apigenin consumption results in the prevention of infection caused by hepatitis C and foot and mouth disease virus at the post-entry stage (Figure 1). Reactivation of Epstein-Barr virus (EBV) and the development of virions by EBV-positive neural progenitor cells (NPC) are both inhibited by apigenin [42]. The active fraction (Fc) of Sambucus gaudichaudiana having apigenin as a major component could exhibit antiviral activity against poliovirus within 4 h [39,43]. Antiviral activities of apigenin were reported against the hepatitis C virus (HCV) by host factor modulation resulting in the reduction of the production of miR122, which facilitates the spread of HCV infection in vitro. Apigenin has also been used for the screening of antiviral activity against the newly developed enterovirus (EV-A71, Fuyang and BrCr strains). It inhibits enterovirus infection by blocking the virus entry sites (Figure 1). It can also disturb the interference between the viral genome with the heterogeneous nuclear ribonucleoprotein A1 and A2 (hnRNP) [44].
Furthermore, apigenin can also suppress cellular apoptosis through the cleavage of caspase-3, which is thought to be a key step in the release of viral progeny. It is also recognized as an ROS scavenger and can reduce the infection caused by ROS in EV-A71 damaged cells. Moreover, it can reduce infection-related cytokine levels. Apigenin was found to decrease the activity of the C and N junction kinase pathway, which is necessary for enterovirus replication. Despite the antiviral inhibitory activity of apigenin, some viruses including Coxsackie virus A16 (CAV16) are unaffected by apigenin consumption. EV71 is a member of the Picornaviridae and belongs to the enterovirus genus. It causes hand, foot and mouth disease (HFMD). High apigenin consumption may inhibit the replication of EV71 and EV71-mediated cytotoxic effects. The expression of viral polyproteins and the up-regulation of several cytokine-related pathways are inhibited by apigenin intake [45]. Studies revealed that the major ingredient of Ocimum basilicum (commonly called sweet basil) is apigenin, which shows antiviral potential [46].
In domestic pigs, the African swine fever virus (ASFV) can cause severe illness. In vitro analysis showed that apigenin has a dose-dependent anti-ASFV activity. Apigenin given to vero cells for one hour after infection reduced ASFV by more than 3 logs. Apigenin can also hinder the synthesis of ASFV-associated proteins as well as the formation of viral factories. Continuous apigenin therapy can reduce cytopathic effects in ASFV-infected cells [47]. Downstream inhibition of MAP kinase signal transduction is facilitated by apigenin consumption which ultimately prevents viral infection. Apigenin has strong antiviral activity against the C4 strain of the enterovirus genotype. At a 24.74 micromolar concentration, it decreases the cytopathogenic effect in human embryonic 293S cells by 50%. In addition, apigenin inhibits the translation of viral mRNA [43]. A compound named 7-O D-glucopyranose derived from the natural herb has the ability to protect T-cell lines from HIV [48,49].

5. Hypoglycemic Effects of Apigenin

Depending on the severity and scope of the condition, hypoglycemia can lead to unconsciousness and even death. It can produce ischemia or depolarization/repolarization-related alterations in the heart which can result in sudden cardiac arrest. Impaired cognitive behavior, especially in young children, causes long-term and potentially detrimental repercussions on their intellectual functionality [50]. A study has found that apigenin compound 6-C-(2″-O-α-l-rhamnopyranosyl)-β-l-fucopyranoside) can effectively lower the C6H12O6 level in the blood and activate the secretion of insulin [51] (Figure 1). Other studies have shown that the 6-C-β-l-fucopyranoside compound can increase glycogen synthesis in host muscle fibers [52] (Figure 1). Apigenin is found in celery vegetables that affect the metabolism of glucose and transmission in peripheral tissues [53]. The leaves of new Bouldia laevis plant have antidiabetic activity against patients having diabetes mellitus (metabolic endocrine disorder) [54].
Apigenin has been shown to suppress the activity of liver gluconeogenic enzymes, which increases the metabolic rate of glucose. It also regulates insulin secretion and produces a strong protective effect on injured pancreatic cells after streptozotocin (STZ) induction [55]. A case study reported in January 2021 found hypoglycemic effects of apigenin in diabetic rats. The sugar level and body weight of rats declined over days due to apigenin consumption [56]. Vitexin compounds in apigenin specifically reduced the blood sugar level and body weight of rats [57]. Apigenin reduced diabetic symptomology by suppressing the MAPK signaling activation, which in turn suppressed the caspase 3 and nitric oxide (NO) signaling pathway axis. Kaempferol (another flavonoid) has also been proven to reduce T2DM symptomology and its complications [57].

6. Role of Apigenin on Biological System

6.1. Apoptosis

Apoptosis can be regulated when promoters and inhibitors help to activate apoptotic processes, several genes, and inhibitory factors [58]. It can also be triggered when apoptotic receptors are exposed to their ligands or cytochrome C liberated into the cytosol [59].
IAP (inhibitors of apoptosis) family members have the potential to inhibit the function of specific caspases and induce apoptosis in them [60]. Due to IAP overexpression, several IAP inhibitors including IAP-1 and IAP-2 were also observed in prostate cancer cell lines [61]. Studies suggested that when Bax (pro-apoptotic protein) interacts with the C-terminus of Ku70 protein, the normal apoptotic process can be inhibited [62]. Recent studies have found that the Ku70 protein is responsible for cancer progression and metastasis; however, Ku70–Bax bioconjugate can also act as a promising therapeutic target [63]. Apigenin consumption may drastically decrease the total number of tumor cells and activate the process of apoptosis leading to a reduction in Bcl-xL and Bcl-2 levels and activating the Bax protein. The expression of HDAC1 gene and class I histone deacetylases can also be inhibited by apigenin therapy, which causes apoptosis in a cancer cell. Apigenin also decreased HDAC1 occupancy at the XIAP promoter, implying that histone deacetylation is essential for XIAP downregulation. These findings have confirmed the apoptotic behavior of apigenin in prostate cancer cells [64].
Apoptosis-associated proteins in apoptotic cells have been the subject of discussion in many studies. Studies have suggested that apigenin can be treated against various types of cancers [8]. Apigenin causes apoptosis more efficiently in blood cancer cells as compared to other cancer cells [65] however, the anticancer activity of apigenin in the cholangiocarcinoma cell lines has not been discovered yet [66]. Apigenin blocks the PI3K/AKT pathway, and induces several caspases and cell cycle arrest in different cancer types [67].
Apigenin can also inhibit AKT signal transduction. AKT is an anti-apoptotic signaling molecule that can mediate PI3K-dependent cell survival responses [68]. Bcl-2 proteins are significantly involved in apoptosis. The mitochondrial apoptotic pathway is activated because of the changes that occurred in the pro- and anti-apoptotic behavior of these proteins [69]. Apigenin treatment induced apoptosis in T24 cells and suppressed AKT phosphorylation in a dose-dependent manner and ultimately blocked the PI3K/AKT pathway, which inhibits cancer proliferation.
A study reported that apigenin treatment on the HeLa cell lines can remarkably enhance the activity of p21/WAF1 protein. The transcriptional upregulation of this protein is p53-dependent. Apigenin also enhances the activity of the Fas/APO-1 and caspase-3 proteins, and both proteins are directly linked to apoptosis. The results of this study showed that apigenin can act as a cervical cancer inhibitory agent [70]. Apigenin shows synergistic effects in combined treatment with sorafenib in the HepG2 human cell line (HCC) in less time and statistically reduces the viability of tumor cells, migration, gene expression and apoptosis. Anti-cancerous drugs are effective but they also harm living bodies due to their toxicities. A combination of these anti-cancerous drugs with apigenin has shown promising potential against tumors. RT-PCR results demonstrated that the efficacy of apigenin showed a notable rise in the expression of p21 protein in the 50 uM 4, 5, 7-Trihydroxyflavone treated group, which indicated that apigenin has strong anti-cancer activity on human cancer cells [71]. Apigenin can induce autophagy and apoptosis, which inhibits cancer metastasis. It can prevent cell mobility, maintain the cell cycle, and stimulate the immune system [72]. In a study, apigenin-treated human lung cancer H460 cell lines were morphologically analyzed at different times to check the pro-apoptotic activity. Results suggested that apigenin can promote apoptosis in H460 cells by decreasing the membrane potential of the mitochondrial-dependent pathway, increasing the expression of ROS and calcium accumulation in lung cancer [73]. This bioactive compound can effectively work against human breast cancer cell lines T47D and MDA-MB-231. The mortality of both cell lines was attributed to apoptosis, high levels of caspase, poly (ADP-ribose) polymerase PARP-cleavage, and a high Bax/Bcl ratio [74]. Furthermore, apigenin-treated cells inhibited autophagy. Such combinations of autophagy induced apoptosis and provided significant potential to combat human breast cancer. It is hypothesized that apigenin consumption promotes apoptosis and activates extrinsic and intrinsic apoptotic pathways. A study reported the potency of apigenin in human osteosarcoma by inducing apoptosis in a human bone osteosarcoma epithelial cell line, (U-2OS) and it also suppresses the xenograft tumor growth [75].

6.2. Immune Response

Plant secondary metabolites (such as alkaloids, phenolics and terpenes) have known medicinal properties including the antioxidant, antidiabetic, antimicrobial, antimutagenic, anti-inflammatory and anti-clotting [76]. Studies have shown that apigenin intake in human hosts significantly fluctuates the immune system response. The immune cells have a specified protein named programmed cell death 1 (PD1); its receptors are PD-L1 (PD-Ligand1) and PD-L2 (PD-Ligand2). These receptors are present on the surface of either macrophages or on the dendritic cells [77]. When PD1 comes in contact with its receptor, the output would be in the form of activation of a PD1 signaling pathway. Several autoimmune diseases/inflammation may occur if PD1/PD-L1 activates the immune system at an inappropriate time [78]. Pardoll et al. reported that PD-L1 complex can be expressed in cancerous cells, leading to antigenic escape in them [79].
Apigenin can stimulate the expression of the NF-κB signaling cascade, which enhances the level of WBCs in the human body [80]. Another study on the rat model reported that consumption of soluble apigenin (Apigenin K) may not only lower the inflammatory and colonic damage level in rats but also stabilize the activity of TNF-α, IL-6 and chemokine ligands. Li et al. suggested that apigenin intake can also suppress ovalbumin (OVA) activity in Th17 cells and ameliorates asthmatic progression in the white blood cells of mice [81].
Cancerous cells can also cause immunity damage in the host body by the inhibition of T effector cell expression and also enhance the levels of T regulatory cells. Research on the pancreatic mouse model reported that apigenin effectively enhanced the activity of T cells and downregulated the activity of T regulatory cells, which increased the survival rate in mice [82]. Continuous consumption of apigenin for two weeks could decrease the level of IgE antibodies, while no apparent effect has been shown on IgG, IgM and IgA antibodies levels [83].

6.3. Autophagy

Autophagy is associated with lipid breakdown and helps in the metabolism of lipid droplets (LDs). However, no investigation has been done to evaluate apigenin’s lipid-lowering mechanism in the context of autophagy. Apigenin has previously been found to have anti-adipogenic effects in HepG2 cells. Apigenin’s role in autophagy and lipid accumulation in palmitic acid (PA)-induced HepG2 cells has also been documented in the literature. Apigenin increased autophagosome formation and the cytosolic light chain ratio (LC3-II/I) in HepG2 cells. In PA-treated HepG2 cells, apigenin promoted autophagy and enhanced autophagic lipid breakdown [84,85]. Apigenin helped to increase the metabolism of lipids by inducing the AMPK-SREBP pathway in HepG2 cells. In addition, apigenin decreased cell proliferation and activated autophagy by inhibiting different cellular pathways in HepG2 cells. p-mTOR, P62, and LC3 expression in PA-treated HepG2 cells were investigated by western blotting. These cells were treated with apigenin and PA at various concentrations over a variety of periods, and it was observed that the cytosolic light chain ratio in apigenin exposed cells increased after every 3 h [86].
Apigenin therapy may alleviate the restricted autophagic flow and lipid buildup caused by PA through a new mechanism. Furthermore, it was demonstrated that the autophagic mechanism is critical for lipid breakdown. Both PA and apigenin increased autophagosome accumulation, and the cytosolic light chain ratio was enhanced even more. A recent study discovered that hepatic steatosis occurred due to the deficiency of autophagic flux in the host body [87]. Excessive lipid aggregation has been demonstrated to hinder autophagy by reducing lysosomal acidification and the activation of different hydrolyses in cells. It is worth noting that conventional medicines such as metformin are useful for the treatment of metabolic illnesses, which helps to reduce the hepatic fat levels via autophagy [88].
Similarly, according to a prior study, epigallocatechin gallate and kaempferol improved the breakdown of LDs by facilitating autophagic flow [89]. According to a lactate dehydrogenase (LDH) release test, apigenin administration promoted cell mortality in primary human epidermal keratinocytes (HEKs) and the cutaneous squamous cell carcinoma cell line COLO-16. Apigenin therapy, on the other hand, restored autophagy inhibition in ultraviolet B treated HEKs. Furthermore, apigenin consumption reversed the ultraviolet B radiation-induced downregulation of the ataxia-telangiectasia and the unfolded protein response (UPR) regulatory proteins in HEK cells. UVB-induced apoptosis and cell death in HEKs were likewise reduced by apigenin administration. It was demonstrated that apigenin has a unique effect in UVB-damaged keratinocytes, implying that it could be used as a photoprotective agent [90]. It can cause apoptosis in PEL cells and reduce the level of intracellular ROS [91].
Apigenin treatment for two weeks dramatically reduced mTOR activity. The results of apigenin administration against ULK1 and AMPK pathways were opposite to that of mTOR signaling transduction, which regulates autophagy via mTOR/AMPK/ULK1 pathway. Apigenin has antidepressant effects in both chronic restraint stress rats and mice [92]. Autophagy is considerably impaired in the sera of people with severe depression in clinical studies. It has been linked to the development of several neurological illnesses such as Alzheimer’s disease, Parkinson’s disease as well as depression [93]. The current study highlighted some of the polyphenolic and bioactive potential of apigenin. However, more metabonomic, proteomic, and transcriptomic research is needed as the next step in developing apigenin as a therapeutic drug.

7. Role of Apigenin in Cell Signaling Pathways

7.1. PI3K/AKT/MTOR Pathway

The phosphatidylinositol 3-kinase, protein kinase B and the mammalian target of the rapamycin (PI3K/AKT/mTOR) cascade have been considered to regulate the morphological characteristics of healthy cells and promote the activity of different genes, metastasis, tumorigenesis, transcription, translation, cell proliferation, migration and cell development in cancerous cells [94,95,96]. PI3K belongs to the lipid kinase family and is divided into the following classes: class-I, class-II and class-III. class-I and class-III are heterodimers; the former has p85 and p110 subunits, while class-III has Vps 34 and Vps15/p150 subunits. Class-I is not only a part of the insulin signal transduction but also has the potential to regulate the AKT pathway, the immune response and plays a significant role in cell proliferation. Class-III is involved in intracellular trafficking and phagocytosis. Class-II PI3Ks are monomers and have a specialized C-terminal domain that typically binds with lipids instead of Ca+2 because of the absence of aspartic acid residues [97,98,99,100]. The central participant of PI3K/AKT/MTOR is AKT, which is activated by the phosphorylation of PI3K. It can trigger ribosomes for protein synthesis and helps in the suppression of TSC1/2 complex and other cell cycle inhibitors, resulting in the activation of the mTOR pathway [101]. mTOR also belongs to the kinase family, having a molecular weight of about 289 kDa. It helps in phagocytosis, synthesis, transcription and translation. Studies have shown that this transduction signaling cascade is linked with cancer progression [102,103].
On the activation of this pathway, several of its proteins are regulated by the phosphorylation of the AKT molecule, which is responsible for cell growth and survival. Apigenin consumption can significantly halt the ATP binding site of phosphatidylinositol 3-kinase, thereby ultimately reducing the functionality of AKT [104,105] (Figure 2). The expression of a transcription factor named FOXO3a (Forkhead box O3) is enhanced by apigenin in breast cancer patients, which downregulates AKT signal transduction. The survival of breast cancer cells is inhibited by upregulating the p21, CIP/WAF, KIPI, and p27 inhibitors [106,107] (Figure 2). Researchers have reported that apigenin consumption has the potential to inhibit P70S6K and S6 proteins, particularly in JAR and JEG3 cells, and to upregulate the expression of ERK1/2 as well as P90RSK proteins. Under specific conditions and cancer types, apigenin consumption can also upregulate the expression of ERK1/2 as well as P90RSK. Some inhibitors like LY294002 and U0126 directly inhibit the expression of PI3K and ERK1/2 [107].
Apigenin also has the potential to stabilize the blood glucose concentration in the body by lowering the activity of glucose transporter protein (GLUT-1) (Figure 2). The reduced expression of GLUT-1 is observed by the inhibition of the PI3K/AKT signaling cascade. This flavonoid helps to activate the Bax protein and inhibits the expression of Bcl2. As a result of this, cytochrome C releases, which directly activates the expression of caspase 3 and 9 and leads to apoptosis and anti-proliferative effects (Figure 2). A study on a mice model by Bridgeman et al. suggested that ultraviolet B (UVB) radiation exposure can activate PI3K/AKT/MTOR, which can cause the progression of skin cancer. Interestingly, apigenin suppressed mTOR activity, raised the UVB-induced phagocytosis and reduced cancerous cell proliferation and growth [108] (Figure 2).
Apigenin has demonstrated its efficacy in the treatment of prostate cancer, producing an inhibitor effect that aids in the cancer cell invasion and migration in a dose-dependent manner. In melanoma A375 and C8161 cell lines, 40 mg of apigenin was found to inhibit the activity of cancer cells via AKT/mTOR signal transduction [109].

7.2. JAK/STAT Pathway

JAK/STAT signaling cascade is responsible for apoptosis, cell division, cell proliferation, tumorigenesis, and arbitrates several cellular responses to cytokines [110] (Figure 3). JAK mediates its control by mediating four proteins that include JAK1, JAK2, JAK3, TYK2 and four domains: FERM, SH2, kinase and pseudokinase having almost 400, 100, 250 and 300 residues, respectively [111], while STAT contains a total of seven proteins: STAT1, STAT2, STAT3, STAT4, STAT5A, STAT5B and STAT6, which have 575–680 residues. It also has several transcriptional activation domains located at the C terminal of amino acids [112].
The JAK/STAT cellular pathway is linked with multiple cancers, and it has been investigated in various animal models [113]. The STAT protein is phosphorylated by several kinases that form dimers which are then moved to the nucleus and act there as an activation factor for transcription. The potential role of apigenin was investigated previously; it prevents the growth and development of tumor cells by inhibiting JAK/SRC phosphorylation, resulting in the suppression of STAT3 activation, which further inhibits the transference of STAT dimers to the nucleus [114] (Figure 3). Various genes are regulated by STAT3 including MMPs, TWIST1 and VEGF, and studies have suggested that these genes trigger metastasis, cell migration and tumor formation, but the expression of these genes is inhibited by apigenin treatment, which significantly blocked the activity of STAT3 [115] (Figure 3). Apigenin can also halt STAT3, STAT5 gene expression and enhance the activity of STAT1, STAT2 and LMWPTP, which inhibits cell proliferation, tumor invasion and tumorigenesis [116] (Figure 3).

7.3. NF-kB Pathway

The efficiency of apigenin is regulated by the repression of COX-2, iNOS, MMP-9, and cyclinD1 gene expression. All of these genes are controlled by the Nuclear Factor-kappaB (NF-kB) pathway. Apigenin is also reported to cause apoptosis in many cancer types [117] via the inhibition of DNA replication, protein kinase inhibition, ROS production, damage to mitochondria, and interference with Ku70-Bax interaction. Increased survival and proliferation are reported to be involved with NF-kB signal transduction which induces transcription of those targeted genes with their product, which can inhibit different stages of programmed cell death [118,119].
This pathway is composed of different dimers that are made with several different segments: NF-kB1 (p50/p105); NF-B2 (p52/100); RelA (p65); RelB; and c-Rel proteins. The inhibitors of the IB (Investigator’s Brochure) family includes IB, IkB, Bcl-3, p100, and p105, which control the NF-kB proteins. NF-kB heterodimers having p65 and p50 can be found in the cytoplasm and are bound to IB subunits in their inactive form. The activation of kinase complex (IKK) phosphorylates the IB subunit at a serine residue and induces ubiquitination and proteasomal degradation, which causes localization of the nucleus and the transcriptional activation of NF-kB [120] (Figure 4).
Hematological malignancies, as well as cancer of the skin, uterine cervix, lung, esophagus, prostate and pancreas, have all been connected with abnormal NF-kB activation [121]. Researchers noted that NF-kB is stimulated constitutively in human prostate cancer tissue, prostate tumor xenografts and in the transgenic adenocarcinoma of the mouse prostate model (TRAMP), which mimics progressive types of human prostate cancer [122]. The enhanced activity of NF-kB is connected to the production of different diseases, and its nuclear localization is linked with the biochemical recurrence and poor state of prognosis [123].
In addition, the IB protein family members (IB, IB, IkB, Bcl-3, p100, and p105) control the NF-kB family members. NF-kB exists in the heterodimeric state in the cytoplasm and binds to IB in an inactive state. TNF, FasL, and TRAIL are only a few of the signal molecules that activate the IKK complex, causing IB to be phosphorylated and degraded. NF-kB then flows towards the nucleus. Pro-survival genes, cell cycle-related genes (cyclin D1), (VEGF), inflammatory cytokines, and tumorous genes are among the target genes activated by NF-kB (COX-2) [124]. Apigenin therapy prevents NF-kB activation in most cases, both in vitro and in vivo. Shukla et al. demonstrated that apigenin consumed by transgenic adenocarcinoma prostate mice prevented prostate tumorigenesis by interfering with the NF-kB signaling pathway in a prostate mouse model. Apigenin therapy can also reduce prostate tumor volume and significantly eliminate cancer cells [125]. According to the studies, the use of apigenin can inhibit the phosphorylation and degradation of IB by halting IKK activation, which suppressed the activation of NF-kB [126] (Figure 4).
Apigenin cannot affect NF-kB expression in the human non-small cell lung cancer cell lines (A549), but it can prevent the translocation of NF-kB from the cytoplasm to the nucleus, resulting in the inhibition of apoptosis-blocking target genes. Apigenin also inhibits the degradation of IB in lung cancer, preventing IB from being separated from the NF-kB heterodimer [127,128]. Apigenin therapy inhibited NF-kB nuclear translocation and AKT activation, as well as modulating MAPK signaling pathways, in malignant mesothelioma in vitro and in vivo [129,130]. It is reported that epidermal growth factor receptor (EGFR) activates the PI3K/AKT pathway, which may lead to the activation of IKK, IKBa phosphorylation on serine 32/36, NF-kb translocation and HER-2 signaling through protein kinase (Ck-2), which results in the NF-kb transactivational activity (Figure 4). Apigenin also blocks LPS and miR-33 activity and ultimately inhibits the expression of several cytokines and tumor causing genes in the nucleus. As a result of this, it can prevent cancer cell proliferation and metastasis (Figure 4).

7.4. ERK/MAPK Pathway

Apigenin has been shown to activate the cleaved caspase-3 and PARP expression sites, down-regulate the Twist1, p-mTOR, ERK1/2 proteins, deactivate FAK/ERK1/2 pathways and halt the phosphorylation of STAT3 in different melanoma cases [131,132] (Figure 5). The ERK family is organized into three different key categories which include ERK/MAPK, p38 kinase and c-Jun/SAPK. Cell development and homeostasis are dependent on these kinases. Several members of the ERK protein family take part to regulate the MAPK-ERK pathway. The proper and stable functioning of both the upstream and downstream targets of the MAPK/ERK pathway is disrupted by the overexpression of any family member of the ERK family [133]. ERK signaling is stimulated in a variety of cancers by mutations in kinase genes (Figure 5). Apigenin is thought to influence the ERK signaling cascade. It can inhibit the MAPK/ERK potential of proliferation by modulating the expression of AKT, ERK, and NF-kB [134] (Figure 5).
Apigenin can stimulate AKT and ERK inhibition because of the impact of the ABT-263 molecule, inducing apoptosis, as observed in in vitro and in vivo experiments [135] (Figure 5). It has shown the capability to trigger apoptosis in prostate cancer cells. In a prostate cancer mouse model, it directly attacks the IGF/IGFBP-3 protein and lowers its expression by reducing p-AKT and ERK1/2 [136].
Furthermore, apigenin is linked with the inhibition of proliferation in cutaneous melanoma cells of humans. It also inhibits ERK1/2 and FAK phosphorylation, which leads to decreased integrin synthesis [130,137] (Figure 5). A study reported that apigenin can increase the levels of ERK1/2 and inhibit the activation of p38 kinase in lung and prostate cancer cell lines. Results showed that apigenin can induce apoptosis by inhibiting the MAPK/ERK signaling pathway [138,139].
To block the AKT signaling pathway, apigenin has been shown to affect the MAPK/ERK signaling cascade in several malignancies. Apigenin substantially inhibits cell proliferation, invasion, migration, and triggered G2/M phase arrest and death in human melanoma A375 and C8161 cell lines, and reduces the activity of p-ERK1/2, p-AKT, and p-mTOR [140]. It can increase TRAIL-induced apoptosis in non-small cell lung cancer cells by inducing DR4/DR5, AKT, ERK, and NF-B signaling. Apigenin was suggested to improve ABT-263-induced anticancerous activity in the HCT116 and DLD1 cell lines in in vitro and in vivo by inhibiting the pro-survival regulators AKT and ERK [132]. Furthermore, in an autochthonous mouse prostate cancer model, the co-inhibition of AKT and ERK signaling was reported. Apigenin treatment successfully slowed the progression of prostate cancer in mouse models by lowering IGF/IGFBP-3 and reducing p-AKT and p-ERK1/2 expression [141]. In other investigations, this flavonoid inhibited the ERK signaling pathway as well as protein kinases including focal adhesion kinase (FAK) (Figure 5). Hasnat et al. reported that apigenin produced anoikis (programmed cell death) in human cutaneous melanoma cells by lowering integrin protein levels and suppressing FAK and ERK1/2 phosphorylation [139] (Figure 5). Apigenin also inhibited the effects of 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone on pancreatic cancer cell proliferation and their migration in the pancreatic cancerous cells by targeting the protein kinases. i.e., FAK and ERK. Shukla et al. found that the administration of apigenin raised phosphorylation of ERK1/2 and JNK1/2 while decreasing the phosphorylation of p38 in human prostate cancer LNCaP and PC-3 cells. The regulation of this flavonoid in the MAPK pathway resulted in apigenin-induced cell cycle arrest in the G0/G1 phase. Apigenin was discovered to cause the death of cancer cells and diminish cancer cell survival in human choriocarcinoma cells by inhibiting the AKT-mTOR pathway and by boosting the phosphorylation of ERK1/2 and P90RSK in an appropriate dose-dependent manner [107]. Apigenin can directly inhibit the activity of membrane receptors and suppresses the expression of protooncogenes in the nucleus. It can also block ERK and JNK activity and downregulate Atrogin-1 gene expression (Figure 5).

8. Possible Side Effects of Apigenin

In general, intentionally consuming higher doses of dietary flavonoids, such as apigenin, is considered safe and may even have health benefits. There is a low risk of toxicity. The amount of apigenin in an individual’s diet is highly unlikely to reach an amount that can cause harm [142].
There is a slightly higher risk of side effects when a person intentionally takes higher doses of a dietary supplement. The potential side effects of apigenin may include:
  • Upset stomach
  • Muscle relaxation
  • Sedation
If a person experiences stomach discomfort after consuming chamomile extract, which is sometimes taken for its high levels of apigenin, then they should discontinue its use immediately. Topicals containing the nutrient may cause skin irritations in some people. A person should contact their healthcare provider before taking apigenin supplements if they take any prescription medications. There is a significant chance of drug interactions with apigenin, particularly if an individual takes cyclosporine, warfarin, or some types of chemotherapy drugs [143].

9. Conclusions

This review has highlighted the biological significance of apigenin and how this bioactive compound acts as a therapeutic agent because of its well-known characteristics including the anti-bacterial, anti-tumor, anti-inflammatory, anti-oxidant, antiviral and as well as hypoglycemic ones. This flavonoid can suppress the activity of different diseases including cancer, those of the cardiovascular system, and diabetes by acting as a therapeutic agent and helping to promote host health. It is hypothesized that apigenin consumption promotes apoptosis in cancer cells and activates extrinsic and intrinsic apoptotic and cancer signaling pathways. Generally, the consumption of higher doses of apigenin is considered safe and may have health benefits. However, there is a significant chance of drug interactions with apigenin, particularly if a person takes cyclosporine, warfarin, or some types of chemotherapy drugs. It is worthy of note that due to the beneficial effects of apigenin, it could be used as a dietary supplement or as a feed additive in the near future. However, more metabonomic, proteomic, and transcriptomic research is needed as the next step in developing apigenin as a therapeutic drug.

Author Contributions

R.A., S.M.S. and R.A.A. conceptualization, data analysis and wrote the first draft of the manuscript; M.K.S., M.I., A.F. and S.G. revised the data and helped in graphical work; M.O.M., M.M. and A.A.K. helped in data collection and drawing figures; B.A. and A.M.E. provided funding and supervision. R.A., W.S. and I.G. critically revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This article received external funding. Research Grant number DSR-2021-01-0369.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

A.M.E. extends his appreciation to the Deanship of Scientific Research at Jouf University for funding his work through Research Grant number DSR-2021-01-0369.

Conflicts of Interest

The authors declare that they have no conflicts of interest.

References

  1. Salehi, B.; Venditti, A.; Sharifi-Rad, M.; Kręgiel, D.; Sharifi-Rad, J.; Durazzo, A.; Lucarini, M.; Santini, A.; Souto, E.B.; Novellino, E.; et al. The therapeutic potential of Apigenin. Int. J. Mol. Sci. 2019, 20, 1305. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Li, A.N.; Li, S.; Zhang, Y.J.; Xu, X.R.; Chen, Y.M.; Li, H. Bin Resources and biological activities of natural polyphenols. Nutrients 2014, 6, 6020–6047. [Google Scholar] [CrossRef] [PubMed]
  3. Bhagwat, S.; Haytowitz, D.B.; Holden, J.M. USDA Database for the Flavonoid Content of Selected Foods Release 3; U.S. Department of Agriculture: Washington, DC, USA, 2011; pp. 1–156.
  4. Jack, C.R., Jr.; Bennett, D.A.; Blennow, K.; Carrillo, M.C.; Dunn, B.; Haeberlein, S.B.; Holtzman, D.M.; Jagust, W.; Jessen, F.; Karlawish, J.; et al. NIA-AA research framework: Toward a biological definition of Alzheimer’s disease. Alzheimer’s Dement. 2018, 14, 535–562. [Google Scholar] [CrossRef] [PubMed]
  5. DeRango-Adem, E.F.; Blay, J. Does Oral Apigenin Have Real Potential for a Therapeutic Effect in the Context of Human Gastrointestinal and Other Cancers? Front. Pharmacol. 2021, 12, 681477. [Google Scholar] [CrossRef] [PubMed]
  6. Wang, M.; Firrman, J.; Liu, L.S.; Yam, K. A review on flavonoid apigenin: Dietary intake, ADME, antimicrobial effects, and interactions with human gut microbiota. Biomed. Res. Int. 2019, 2019, 7010467. [Google Scholar] [CrossRef] [PubMed]
  7. Baumann, L.S. Apigenin. Skin Allergy News 2008, 39, 32. [Google Scholar] [CrossRef]
  8. Kashyap, D.; Sharma, A.; Tuli, H.S.; Sak, K.; Garg, V.K.; Buttar, H.S.; Setzer, W.N.; Sethi, G. Apigenin: A natural bioactive flavone-type molecule with promising therapeutic function. J. Funct. Foods 2018, 48, 457–471. [Google Scholar] [CrossRef]
  9. Hong, S.; Dia, V.P.; Baek, S.J.; Zhong, Q. Nanoencapsulation of apigenin with whey protein isolate: Physicochemical properties, in vitro activity against colorectal cancer cells, and bioavailability. LWT 2022, 154, 112751. [Google Scholar] [CrossRef]
  10. Kim, J.K.; Park, S.U. Letter to the editor: Recent insights into the biological functions of apigenin. EXCLI J. 2020, 19, 984–991. [Google Scholar] [CrossRef]
  11. Gupta, S.; Afaq, F.; Mukhtar, H. Selective growth-inhibitory, cell-cycle deregulatory and apoptotic response of apigenin in normal versus human prostate carcinoma cells. Biochem. Biophys. Res. Commun. 2001, 287, 914–920. [Google Scholar] [CrossRef]
  12. Singh, P.; Mishra, S.K.; Noel, S.; Sharma, S.; Rath, S.K. Acute exposure of apigenin induces hepatotoxicity in Swiss mice. PLoS ONE 2012, 7, e31964. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Hashem, S.; Ali, T.A.; Akhtar, S.; Nisar, S.; Sageena, G.; Ali, S.; Al-Mannai, S.; Therachiyil, L.; Mir, R.; Elfaki, I.; et al. Targeting cancer signaling pathways by natural products: Exploring promising anti-cancer agents. Biomed. Pharmacother. 2022, 150, 113054. [Google Scholar] [CrossRef] [PubMed]
  14. Takagaki, N.; Sowa, Y.; Oki, T.; Nakanishi, R.; Yogosawa, S.; Sakai, T. Apigenin induces cell cycle arrest and p21/WAF1 expression in a p53-independent pathway. Int. J. Oncol. 2005, 26, 185–189. [Google Scholar] [CrossRef] [PubMed]
  15. Maggioni, D.; Garavello, W.; Rigolio, R.; Pignataro, L.; Gaini, R.; Nicolini, G. Apigenin impairs oral squamous cell carcinoma growth in vitro inducing cell cycle arrest and apoptosis. Int. J. Oncol. 2013, 43, 1675–1682. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Seo, H.S.; Ku, J.M.; Choi, H.S.; Woo, J.K.; Jang, B.H.; Shin, Y.C.; Ko, S.G. Induction of caspase-dependent apoptosis by apigenin by inhibiting STAT3 signaling in HER2-overexpressing MDA-MB-453 breast cancer cells. Anticancer Res. 2014, 34, 2869–2882. [Google Scholar] [PubMed]
  17. Huang, C.S.; Lii, C.K.; Lin, A.H.; Yeh, Y.W.; Yao, H.T.; Li, C.C.; Wang, T.S.; Chen, H.W. Protection by chrysin, apigenin, and luteolin against oxidative stress is mediated by the Nrf2-dependent up-regulation of heme oxygenase 1 and glutamate cysteine ligase in rat primary hepatocytes. Arch. Toxicol. 2013, 87, 167–178. [Google Scholar] [CrossRef]
  18. Farid, A.; Javed, R.; Hayat, M.; Muzammal, M.; Khan, M.H.; Ismail, S.; Rashid, S.A. Screening of Strobilanthes urticifolia wall.ex kuntze for Antitermite and insecticidal activities. Abasyn J. life Sci. 2021, 4, 40–45. [Google Scholar] [CrossRef]
  19. Rezai-Zadeh, K.; Ehrhart, J.; Bai, Y.; Sanberg, P.R.; Bickford, P.; Tan, J.; Douglas, R.D. Apigenin and luteolin modulate microglial activation via inhibition of STAT1-induced CD40 expression. J. Neuroinflamm. 2008, 5, 41. [Google Scholar] [CrossRef] [Green Version]
  20. Jung, U.J.; Cho, Y.Y.; Choi, M.S. Apigenin ameliorates dyslipidemia, hepatic steatosis and insulin resistance by modulating metabolic and transcriptional profiles in the liver of high-fat diet-induced obese mice. Nutrients 2016, 8, 305. [Google Scholar] [CrossRef] [Green Version]
  21. Escande, C.; Nin, V.; Price, N.L.; Capellini, V.; Gomes, A.P.; Barbosa, M.T.; O’Neil, L.; White, T.A.; Sinclair, D.A.; Chini, E.N. Flavonoid apigenin is an inhibitor of the NAD+ase CD38: Implications for cellular NAD+ metabolism, protein acetylation, and treatment of metabolic syndrome. Diabetes 2013, 62, 1084–1093. [Google Scholar] [CrossRef] [Green Version]
  22. Cardenas, H.; Arango, D.; Nicholas, C.; Duarte, S.; Nuovo, G.J.; He, W.; Voss, O.H.; Gonzalez-Mejia, M.E.; Guttridge, D.C.; Grotewold, E.; et al. Dietary apigenin exerts immune-regulatory activity in vivo by reducing NF-κB activity, halting leukocyte infiltration and restoring normal metabolic function. Int. J. Mol. Sci. 2016, 17, 323. [Google Scholar] [CrossRef] [Green Version]
  23. Seo, H.S.; Jo, J.K.; Ku, J.M.; Choi, H.S.; Choi, Y.K.; Woo, J.K.; Kim, H.I.; Kang, S.Y.; Lee, K.M.; Nam, K.W.; et al. Induction of caspase-dependent extrinsic apoptosis by apigenin through inhibition of signal transducer and activator of transcription 3 (STAT3) signalling in HER2-overexpressing BT-474 breast cancer cells. Biosci. Rep. 2015, 35, e00276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Hu, W.; Wang, X.; Wu, L.; Shen, T.; Ji, L.; Zhao, X.; Si, C.L.; Jiang, Y.; Wang, G. Apigenin-7-O-β-d-glucuronide inhibits LPS-induced inflammation through the inactivation of AP-1 and MAPK signaling pathways in RAW 264.7 macrophages and protects mice against endotoxin shock. Food Funct. 2016, 7, 1002–1013. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, S.; Liu, X.; Sun, C.; Yang, J.; Wang, L.; Liu, J.; Gong, L.; Jing, Y. Apigenin Attenuates Experimental Autoimmune Myocarditis by Modulating Th1/Th2 Cytokine Balance in Mice. Inflammation 2016, 39, 678–686. [Google Scholar] [CrossRef] [PubMed]
  26. Li, R.; Wang, X.; Qin, T.; Qu, R.; Ma, S. Apigenin ameliorates chronic mild stress-induced depressive behavior by inhibiting interleukin-1β production and NLRP3 inflammasome activation in the rat brain. Behav. Brain Res. 2016, 296, 318–325. [Google Scholar] [CrossRef]
  27. Calvo-Guirado, J.L.; López-López, P.J.; Domínguez, M.F.; Gosálvez, M.M.; Prados-Frutos, J.C.; Gehrke, S.A. Histologic evaluation of new bone in post-extraction sockets induced by melatonin and apigenin: An experimental study in American fox hound dogs. Clin. Oral Implants Res. 2018, 29, 1176. [Google Scholar] [CrossRef]
  28. Guerra, J.A.; Molina, M.F.; Abad, M.J.; Villar, A.M.; Paulina, B. Inhibition of inducible nitric oxide synthase and cyclooxygenase-2 expression by flavonoids isolated from Tanacetum microphyllum. Int. Immunopharmacol. 2006, 6, 1723–1728. [Google Scholar] [CrossRef]
  29. Adamczak, A.; Ożarowski, M.; Karpiński, T.M. Antibacterial activity of some flavonoids and organic acids widely distributed in plants. J. Clin. Med. 2020, 9, 109. [Google Scholar] [CrossRef] [Green Version]
  30. Cha, S.; Kim, G.-U.; Cha, J.-D. Synergistic antimicrobial activity of apigenin against oral Pathogens. Int. J. Eng. Res. Sci. 2016, 2, 27–37. [Google Scholar]
  31. Ali, F.; Rahul; Naz, F.; Jyoti, S.; Siddique, Y.H. Health functionality of apigenin: A review. Int. J. Food Prop. 2017, 20, 1197–1238. [Google Scholar] [CrossRef]
  32. Nayaka, H.B.; Londonkar, R.L.; Umesh, M.K.; Tukappa, A. Antibacterial Attributes of Apigenin, Isolated from Portulaca oleracea L. Int. J. Bacteriol. 2014, 2014, 175851. [Google Scholar] [CrossRef] [Green Version]
  33. Eumkeb, G.; Chukrathok, S. Synergistic activity and mechanism of action of ceftazidime and apigenin combination against ceftazidime-resistant Enterobacter cloacae. Phytomedicine 2013, 20, 262–269. [Google Scholar] [CrossRef] [PubMed]
  34. Banerjee, K.; Banerjee, S.; Das, S.; Mandal, M. Probing the potential of apigenin liposomes in enhancing bacterial membrane perturbation and integrity loss. J. Colloid Interface Sci. 2015, 453, 48–59. [Google Scholar] [CrossRef] [PubMed]
  35. Koo, H.; Schobel, B.; Scott-Anne, K.; Watson, G.; Bowen, W.H.; Cury, J.A.; Rosalen, P.L.; Park, Y.K. Apigenin and tt-farnesol with fluoride effects on S. mutans biofilms and dental caries. J. Dent. Res. 2005, 84, 1016–1020. [Google Scholar] [CrossRef] [PubMed]
  36. Engelmann, C.; Blot, E.; Panis, Y.; Bauer, S.; Trochon, V.; Nagy, H.J.; Lu, H.; Soria, C. Apigenin—Strong cytostatic and anti-angiogenic action in vitro contrasted by lack of efficacy in vivo. Phytomedicine 2002, 9, 489–495. [Google Scholar] [CrossRef] [PubMed]
  37. Fang, J.; Bao, Y.-Y.; Zhou, S.H.; Fan, J. Apigenin inhibits the proliferation of adenoid cystic carcinoma via suppression of glucose transporter-1. Mol. Med. Rep. 2015, 12, 6461–6466. [Google Scholar] [CrossRef] [Green Version]
  38. Lee, W.J.; Chen, W.K.; Wang, C.J.; Lin, W.L.; Tseng, T.H. Apigenin inhibits HGF-promoted invasive growth and metastasis involving blocking PI3K/Akt pathway and β4 integrin function in MDA-MB-231 breast cancer cells. Toxicol. Appl. Pharmacol. 2008, 226, 178–191. [Google Scholar] [CrossRef]
  39. Stapleton, P.D.; Shah, S.; Ehlert, K.; Hara, Y.; Taylor, P.W. The β-lactam-resistance modifier (-)-epicatechin gallate alters the architecture of the cell wall of Staphylococcus aureas. Microbiology 2007, 153, 2093–2103. [Google Scholar] [CrossRef] [Green Version]
  40. Visintini Jaime, M.F.; Redko, F.; Muschietti, L.V.; Campos, R.H.; Martino, V.S.; Cavallaro, L.V. In vitro antiviral activity of plant extracts from Asteraceae medicinal plants. Virol. J. 2013, 10, 245. [Google Scholar] [CrossRef] [Green Version]
  41. Lalani, S.; Poh, C.L. Flavonoids as antiviral agents for enterovirus A71 (EV-A71). Viruses 2020, 12, 184. [Google Scholar] [CrossRef] [Green Version]
  42. Shibata, C.; Ohno, M.; Otsuka, M.; Kishikawa, T.; Goto, K.; Muroyama, R.; Kato, N.; Yoshikawa, T.; Takata, A.; Koike, K. The flavonoid apigenin inhibits hepatitis C virus replication by decreasing mature microRNA122 levels. Virology 2014, 462–463, 42–48. [Google Scholar] [CrossRef] [PubMed]
  43. Wang, L.; Song, J.; Liu, A.; Xiao, B.; Li, S.; Wen, Z.; Lu, Y.; Du, G. Research Progress of the Antiviral Bioactivities of Natural Flavonoids. Nat. Prod. Bioprospect. 2020, 10, 271–283. [Google Scholar] [CrossRef] [PubMed]
  44. Dai, W.; Bi, J.; Li, F.; Wang, S.; Huang, X.; Meng, X.; Sun, B.; Wang, D.; Kong, W.; Jiang, C.; et al. Antiviral efficacy of flavonoids against enterovirus 71 infection in vitro and in newborn mice. Viruses 2019, 11, 625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Zhang, W.; Qiao, H.; Lv, Y.; Wang, J.; Chen, X.; Hou, Y.; Tan, R.; Li, E. Apigenin inhibits enterovirus-71 infection by disrupting viral RNA association with trans-acting factors. PLoS ONE 2014, 9, e110429. [Google Scholar] [CrossRef]
  46. Lv, X.; Qiu, M.; Chen, D.; Zheng, N.; Jin, Y.; Wu, Z. Apigenin inhibits enterovirus 71 replication through suppressing viral IRES activity and modulating cellular JNK pathway. Antivir. Res. 2014, 109, 30–41. [Google Scholar] [CrossRef]
  47. Chiang, L.C.; Ng, L.T.; Cheng, P.W.; Chiang, W.; Lin, C.C. Antiviral activities of extracts and selected pure constituents of Ocimum basilicum. Clin. Exp. Pharmacol. Physiol. 2005, 32, 811–816. [Google Scholar] [CrossRef]
  48. Hakobyan, A.; Arabyan, E.; Avetisyan, A.; Abroyan, L.; Hakobyan, L.; Zakaryan, H. Apigenin inhibits African swine fever virus infection in vitro. Arch. Virol. 2016, 161, 3445–3453. [Google Scholar] [CrossRef]
  49. Kulshreshtha, A.; Piplani, P. Current pharmacotherapy and putative disease-modifying therapy for Alzheimer’s disease. Neurol. Sci. 2016, 37, 1403–1435. [Google Scholar] [CrossRef]
  50. Lee, J.S.; Kim, H.J.; Lee, Y.S. A new Anti-HIV Flavonoid Glucuronide from Chrysanthemum morifolium. Planta Med. 2003, 69, 859–861. [Google Scholar] [CrossRef]
  51. Kalra, S.; Mukherjee, J.; Venkataraman, S.; Bantwal, G.; Shaikh, S.; Saboo, B.; Das, A.; Ramachandran, A. Hypoglycemia: The neglected complication. Indian J. Endocrinol. Metab. 2013, 17, 819. [Google Scholar] [CrossRef]
  52. Cazarolli, L.H.; Folador, P.; Moresco, H.H.; Brighente, I.M.C.; Pizzolatti, M.G.; Silva, F.R.M.B. Mechanism of action of the stimulatory effect of apigenin-6-C-(2″-O-α-l-rhamnopyranosyl)-β-l-fucopyranoside on 14C-glucose uptake. Chem. Biol. Interact. 2009, 179, 407–412. [Google Scholar] [CrossRef] [PubMed]
  53. Cazarolli, L.H.; Folador, P.; Moresco, H.H.; Brighente, I.M.C.; Pizzolatti, M.G.; Silva, F.R.M.B. Stimulatory effect of apigenin-6-C-β-l-fucopyranoside on insulin secretion and glycogen synthesis. Eur. J. Med. Chem. 2009, 44, 4668–4673. [Google Scholar] [CrossRef] [PubMed]
  54. Ragab El Barky, A.; Abdel hamid Ezz, A.; Mostafa Mohammed, T. The Potential Role of apigenin in Diabetes Mellitus Stem. Int. J. Clin. Case Rep. Rev. 2020, 3, 1–3. [Google Scholar] [CrossRef]
  55. Osigwe, C.C.; Akah, P.A.; Nworu, C.S.; Okoye, F.B.C. Apigenin: A methanol fraction component of Newbouldia laevis leaf, as a potential antidiabetic agent. J. Phytopharm. 2017, 6, 38–44. [Google Scholar] [CrossRef]
  56. Anandan, S.; Urooj, A. Hypoglycemic effects of apigenin from Morus indica in streptozotocin induced diabetic rats. Int. J. Curr. Res. Rev. 2021, 13, 100–105. [Google Scholar] [CrossRef]
  57. Abdel-Megeed, R.M.; El Newary, S.A.; Kadry, M.O.; Ghanem, H.Z.; El-Shesheny, R.A.; Said-Al Ahl, H.A.H.; Abdel-Hamid, A.H.Z. Hyssopus officinalis exerts hypoglycemic effects on streptozotocin-induced diabetic rats via modulating GSK-3β, C-fos, NF-κB, ABCA1 and ABGA1 gene expression. J. Diabetes Metab. Disord. 2020, 19, 483–491. [Google Scholar] [CrossRef]
  58. Abd El-Ghffar, E.A.; Hegazi, N.M.; Saad, H.H.; Soliman, M.M.; El-Raey, M.A.; Shehata, S.M.; Barakat, A.; Yasri, A.; Sobeh, M. HPLC-ESI-MS/MS analysis of beet (Beta vulgaris) leaves and its beneficial properties in type 1 diabetic rats. Biomed. Pharmacother. 2019, 120, 109541. [Google Scholar] [CrossRef]
  59. Deveraux, Q.L.; Schendel, S.L.; Reed, J.C. Antiapoptotic proteins: The Bcl-2 and inhibitor of apoptosis protein families. Cardiol. Clin. 2001, 19, 57–74. [Google Scholar] [CrossRef]
  60. Gupta, K.; Thakur, V.S.; Bhaskaran, N.; Nawab, A.; Babcook, M.A.; Jackson, M.W.; Gupta, S. Green Tea Polyphenols Induce p53-Dependent and p53-Independent Apoptosis in Prostate Cancer Cells through Two Distinct Mechanisms. PLoS ONE 2012, 7, e52572. [Google Scholar] [CrossRef]
  61. Dubrez-Daloz, L.; Dupoux, A.; Cartier, J. IAPs: More than just inhibitors of apoptosis proteins. Cell Cycle 2008, 7, 1036–1046. [Google Scholar] [CrossRef] [Green Version]
  62. McEleny, K.R.; Watson, R.W.G.; Coffey, R.N.T.; O’Neill, A.J.; Fitzpatrick, J.M. Inhibitors of apoptosis proteins in prostate cancer cell lines. Prostate 2002, 51, 133–140. [Google Scholar] [CrossRef] [PubMed]
  63. Kale, J.; Osterlund, E.J.; Andrews, D.W. BCL-2 family proteins: Changing partners in the dance towards death. Cell Death Differ. 2018, 25, 65–80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Xu, Y.; Xin, Y.; Diao, Y.; Lu, C.; Fu, J.; Luo, L.; Yin, Z. Synergistic effects of apigenin and paclitaxel on apoptosis of cancer cells. PLoS ONE 2011, 6, e29169. [Google Scholar] [CrossRef] [Green Version]
  65. Imran, M.; Aslam Gondal, T.; Atif, M.; Shahbaz, M.; Batool Qaisarani, T.; Hanif Mughal, M.; Salehi, B.; Martorell, M.; Sharifi-Rad, J. Apigenin as an anticancer agent. Phyther. Res. 2020, 34, 1812–1828. [Google Scholar] [CrossRef]
  66. Vargo, M.A.; Voss, O.H.; Poustka, F.; Cardounel, A.J.; Grotewold, E.; Doseff, A.I. Apigenin-induced-apoptosis is mediated by the activation of PKCδ and caspases in leukemia cells. Biochem. Pharmacol. 2006, 72, 681–692. [Google Scholar] [CrossRef]
  67. Subhasitanont, P.; Chokchaichamnankit, D.; Chiablaem, K.; Keeratichamroen, S.; Ngiwsara, L.; Paricharttanakul, N.M.; Lirdprapamongkol, K.; Weeraphan, C.; Svasti, J.; Srisomsap, C. Apigenin inhibits growth and induces apoptosis in human cholangiocarcinoma cells. Oncol. Lett. 2017, 14, 4361–4371. [Google Scholar] [CrossRef] [Green Version]
  68. Zhu, Y.; Mao, Y.; Chen, H.; Lin, Y.; Hu, Z.; Wu, J.; Xu, X.; Xu, X.; Qin, J.; Xie, L. Apigenin promotes apoptosis, inhibits invasion and induces cell cycle arrest of T24 human bladder cancer cells. Cancer Cell Int. 2013, 13, 54. [Google Scholar] [CrossRef] [Green Version]
  69. Franke, T.F.; Hornik, C.P.; Segev, L.; Shostak, G.A.; Sugimoto, C. PI3K/Akt and apoptosis: Size matters. Oncogene 2003, 22, 8983–8998. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Pugazhenthi, S.; Nesterova, A.; Sable, C.; Heidenreich, K.A.; Boxer, L.M.; Heasley, L.E.; Reusch, J.E. Akt/Protein Kinase B Up-Regulates Bcl-2 Expression through cAMP-Response Element-Binding Protein. J. Biol. Chem. 2000, 275, 10761–10766. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Zheng, P.W.; Chiang, L.C.; Lin, C.C. Apigenin induced apoptosis through p53-dependent pathway in human cervical carcinoma cells. Life Sci. 2005, 76, 1367–1379. [Google Scholar] [CrossRef]
  72. Pandey, P.; Khan, F.; Qari, H.A.; Oves, M. Rutin (Bioflavonoid) as cell signaling pathway modulator: Prospects in treatment and chemoprevention. Pharmaceuticals 2021, 14, 1069. [Google Scholar] [CrossRef] [PubMed]
  73. Şirin, N.; Elmas, L.; Seçme, M.; Dodurga, Y. Investigation of possible effects of apigenin, sorafenib and combined applications on apoptosis and cell cycle in hepatocellular cancer cells. Gene 2020, 737, 144428. [Google Scholar] [CrossRef] [PubMed]
  74. Lu, H.F.; Chie, Y.J.; Yang, M.S.; Lu, K.W.; Fu, J.J.; Yang, J.S.; Chen, H.Y.; Hsia, T.C.; Ma, C.Y.; Ip, S.W.; et al. Apigenin induces apoptosis in human lung cancer H460 cells through caspase- and mitochondria-dependent pathways. Hum. Exp. Toxicol. 2011, 30, 1053–1061. [Google Scholar] [CrossRef] [PubMed]
  75. Balez, R.; Steiner, N.; Engel, M.; Muñoz, S.S.; Lum, J.S.; Wu, Y.; Wang, D.; Vallotton, P.; Sachdev, P.; O’Connor, M.; et al. Neuroprotective effects of apigenin against inflammation, neuronal excitability and apoptosis in an induced pluripotent stem cell model of Alzheimer’s disease. Sci. Rep. 2016, 6, 31450. [Google Scholar] [CrossRef] [Green Version]
  76. Lin, C.C.; Chuang, Y.J.; Yu, C.C.; Yang, J.S.; Lu, C.C.; Chiang, J.H.; Lin, J.P.; Tang, N.Y.; Huang, A.C.; Chung, J.G. Apigenin induces apoptosis through mitochondrial dysfunction in U-2 OS human osteosarcoma cells and inhibits osteosarcoma xenograft tumor growth in vivo. J. Agric. Food Chem. 2012, 60, 11395–11402. [Google Scholar] [CrossRef]
  77. Toh, E.Y.S.; Lim, C.L.; Ling, A.P.K.; Chye, S.M.; Koh, R.Y. Overview of the pharmacological activities of aframomum melegueta. Pertanika J. Trop. Agric. Sci. 2019, 42, 1–13. [Google Scholar]
  78. Alexandrescu, D.T.; Ichim, T.E.; Riordan, N.H.; Marincola, F.M.; Di Nardo, A.; Kabigting, F.D.; Dasanu, C.A. Immunotherapy for melanoma: Current status and perspectives. J. Immunother. 2010, 33, 570–590. [Google Scholar] [CrossRef] [Green Version]
  79. Mahoney, K.M.; Rennert, P.D.; Freeman, G.J. Combination cancer immunotherapy and new immunomodulatory targets. Nat. Rev. Drug Discov. 2015, 14, 561–584. [Google Scholar] [CrossRef]
  80. Pardoll, D.M. The blockade of immune checkpoints in cancer immunotherapy. Nat. Rev. Cancer 2012, 12, 252–264. [Google Scholar] [CrossRef] [Green Version]
  81. Mascaraque, C.; González, R.; Suárez, M.D.; Zarzuelo, A.; De Medina, F.S.; Martínez-Augustin, O. Intestinal anti-inflammatory activity of apigenin K in two rat colitis models induced by trinitrobenzenesulfonic acid and dextran sulphate sodium. Br. J. Nutr. 2015, 113, 618–626. [Google Scholar] [CrossRef] [Green Version]
  82. Li, J.; Zhang, B. Apigenin protects ovalbumin-induced asthma through the regulation of Th17 cells. Fitoterapia 2013, 91, 298–304. [Google Scholar] [CrossRef] [PubMed]
  83. Mishra, A.K.; Kadoishi, T.; Wang, X.; Driver, E.; Chen, Z.; Wang, X.J.; Wang, J.H. Squamous cell carcinomas escape immune surveillance via inducing chronic activation and exhaustion of CD8+ T Cells co-expressing PD-1 and LAG-3 inhibitory receptors. Oncotarget 2016, 7, 81341–81356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Yano, S.; Umeda, D.; Maeda, N.; Fujimura, Y.; Yamada, K.; Tachibana, H. Dietary apigenin suppresses IgE and inflammatory cytokines production in C57BL/6N mice. J. Agric. Food Chem. 2006, 54, 5203–5207. [Google Scholar] [CrossRef] [PubMed]
  85. Kwanten, W.J.; Martinet, W.; Michielsen, P.P.; Francque, S.M. Role of autophagy in the pathophysiology of nonalcoholic fatty liver disease: A controversial issue. World J. Gastroenterol. 2014, 20, 7325–7338. [Google Scholar] [CrossRef] [PubMed]
  86. Lu, J.; Meng, Z.; Chen, Y.; Yu, L.; Gao, B.; Zheng, Y.; Guan, S. Apigenin induced autophagy and stimulated autophagic lipid degradation. Food Funct. 2020, 11, 9208–9215. [Google Scholar] [CrossRef]
  87. Wang, Y.; Hao, C.L.; Zhang, Z.H.; Wang, L.H.; Yan, L.N.; Zhang, R.J.; Lin, L.; Yang, Y. Valproic Acid Increased Autophagic Flux in Human Multiple Myeloma Cells In Vitro. Biomed. Pharmacother. 2020, 127, 110167. [Google Scholar] [CrossRef]
  88. Madrigal-Matute, J.; Cuervo, A.M. Regulation of Liver Metabolism by Autophagy. Gastroenterology 2016, 150, 328–339. [Google Scholar] [CrossRef] [Green Version]
  89. Singh, R.; Kaushik, S.; Wang, Y.; Xiang, Y.; Novak, I.; Komatsu, M.; Tanaka, K.; Cuervo, A.M.; Czaja, M.J. Autophagy regulates lipid metabolism. Nature 2009, 458, 1131–1135. [Google Scholar] [CrossRef] [Green Version]
  90. Kim, H.S.; Montana, V.; Jang, H.J.; Parpura, V.; Kim, J.A. Epigallocatechin gallate (EGCG) stimulates autophagy in vascular endothelial cells: A potential role for reducing lipid accumulation. J. Biol. Chem. 2013, 288, 22693–22705. [Google Scholar] [CrossRef] [Green Version]
  91. Li, L.; Li, M.; Xu, S.; Chen, H.; Chen, X.; Gu, H. Apigenin restores impairment of autophagy and downregulation of unfolded protein response regulatory proteins in keratinocytes exposed to ultraviolet B radiation. J. Photochem. Photobiol. B Biol. 2019, 194, 84–95. [Google Scholar] [CrossRef]
  92. Granato, M.; Gilardini Montani, M.S.; Santarelli, R.; D’Orazi, G.; Faggioni, A.; Cirone, M. Apigenin, by activating p53 and inhibiting STAT3, modulates the balance between pro-apoptotic and pro-survival pathways to induce PEL cell death. J. Exp. Clin. Cancer Res. 2017, 36, 167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Zhang, X.; Bu, H.; Jiang, Y.; Sun, G.; Jiang, R.; Huang, X.; Duan, H.; Huang, Z.; Wu, Q. The antidepressant effects of apigenin are associated with the promotion of autophagy via the mTOR/AMPK/ULK1 pathway. Mol. Med. Rep. 2019, 20, 2867–2874. [Google Scholar] [CrossRef] [PubMed]
  94. Gassen, N.C.; Hartmann, J.; Zschocke, J.; Stepan, J.; Hafner, K.; Zellner, A.; Kirmeier, T.; Kollmannsberger, L.; Wagner, K.V.; Dedic, N.; et al. Association of FKBP51 with Priming of Autophagy Pathways and Mediation of Antidepressant Treatment Response: Evidence in Cells, Mice, and Humans. PLoS Med. 2014, 11, e1001755. [Google Scholar] [CrossRef] [PubMed]
  95. 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. 2022, 80, 1–17. [Google Scholar] [CrossRef] [PubMed]
  96. Yu, J.S.L.; Cui, W. Proliferation, survival and metabolism: The role of PI3K/AKT/mTOR signalling in pluripotency and cell fate determination. Development 2016, 143, 3050–3060. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Pavlidou, A.; Vlahos, N.F. Molecular alterations of PI3K/Akt/mTOR pathway: A therapeutic target in endometrial cancer. Sci. World J. 2014, 2014, 709736. [Google Scholar] [CrossRef] [Green Version]
  98. Cantley, L.C. The phosphoinositide 3-kinase pathway. Science 2002, 296, 1655–1657. [Google Scholar] [CrossRef]
  99. Engelman, J.A.; Luo, J.; Cantley, L.C. The evolution of phosphatidylinositol 3-kinases as regulators of growth and metabolism. Nat. Rev. Genet. 2006, 7, 606–619. [Google Scholar] [CrossRef]
  100. Markman, B.; Dienstmann, R.; Tabernero, J. Targeting the PI3K/Akt/mTOR pathway—Beyond rapalogs. Oncotarget 2010, 1, 530–543. [Google Scholar] [CrossRef] [Green Version]
  101. Carpenter, C.L.; Duckworth, B.C.; Auger, K.R.; Cohen, B.; Schaffhausen, B.S.; Cantley, L.C. Purification and characterization of phosphoinositide 3-kinase from rat liver. J. Biol. Chem. 1990, 265, 19704–19711. [Google Scholar] [CrossRef]
  102. Huang, J.; Lam, G.Y.; Brumell, J.H. Autophagy signaling through reactive oxygen species. Antioxid. Redox Signal. 2011, 14, 2215–2231. [Google Scholar] [CrossRef] [PubMed]
  103. Lipton, J.O.; Sahin, M. The Neurology of mTOR. Neuron 2014, 84, 275–291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Hay, N.; Sonenberg, N. Upstream and downstream of mTOR. Genes Dev. 2004, 18, 1926–1945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Brown, J.S.; Banerji, U. Maximising the potential of AKT inhibitors as anti-cancer treatments. Pharmacol. Ther. 2017, 172, 101–115. [Google Scholar] [CrossRef] [PubMed]
  106. Tong, X.; Pelling, J. Targeting the PI3K/Akt/mTOR Axis by Apigenin for Cancer Prevention. Anticancer Agents Med. Chem. 2013, 13, 971–978. [Google Scholar] [CrossRef] [Green Version]
  107. Lin, C.H.; Chang, C.Y.; Lee, K.R.; Lin, H.J.; Chen, T.H.; Wan, L. Flavones inhibit breast cancer proliferation through the Akt/FOXO3a signaling pathway. BMC Cancer 2015, 15, 958. [Google Scholar] [CrossRef] [Green Version]
  108. Lim, W.; Park, S.; Bazer, F.W.; Song, G. Apigenin Reduces Survival of Choriocarcinoma Cells by Inducing Apoptosis via the PI3K/AKT and ERK1/2 MAPK Pathways. J. Cell. Physiol. 2016, 231, 2690–2699. [Google Scholar] [CrossRef]
  109. Bridgeman, B.B.; Wang, P.; Ye, B.; Pelling, J.C.; Volpert, O.V.; Tong, X. Inhibition of mTOR by apigenin in UVB-irradiated keratinocytes: A new implication of skin cancer prevention. Cell. Signal. 2016, 28, 460–468. [Google Scholar] [CrossRef] [Green Version]
  110. Yan, X.; Qi, M.; Li, P.; Zhan, Y.; Shao, H. Apigenin in cancer therapy: Anti-cancer effects and mechanisms of action. Cell Biosci. 2017, 7, 50. [Google Scholar] [CrossRef] [Green Version]
  111. Roberts, S.; Peyman, S.; Speirs, V. Current and Emerging 3D Models to Study Breast Cancer. Adv. Exp. Med. Biol. 2019, 1152, 413–427. [Google Scholar] [CrossRef]
  112. Schindler, C.; Levy, D.E.; Decker, T. JAK-STAT signaling: From interferons to cytokines. J. Biol. Chem. 2007, 282, 20059–20063. [Google Scholar] [CrossRef] [Green Version]
  113. Kiu, H.; Nicholson, S.E. Biology and significance of the JAK/STAT signalling pathways. Growth Factors 2012, 30, 88–106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Xin, P.; Xu, X.; Deng, C.; Liu, S.; Wang, Y.; Zhou, X.; Ma, H.; Wei, D.; Sun, S. The role of JAK/STAT signaling pathway and its inhibitors in diseases. Int. Immunopharmacol. 2020, 80, 106210. [Google Scholar] [CrossRef]
  115. Pérez-Ruiz, E.; Melero, I.; Kopecka, J.; Sarmento-Ribeiro, A.B.; García-Aranda, M.; De Las Rivas, J. Cancer immunotherapy resistance based on immune checkpoints inhibitors: Targets, biomarkers, and remedies. Drug Resist. Updates 2020, 53, 100718. [Google Scholar] [CrossRef]
  116. Cao, H.H.; Chu, J.H.; Kwan, H.Y.; Su, T.; Yu, H.; Cheng, C.Y.; Fu, X.Q.; Guo, H.; Li, T.; Tse, A.K.W.; et al. Inhibition of the STAT3 signaling pathway contributes to apigenin-mediated anti-metastatic effect in melanoma. Sci. Rep. 2016, 6, 21731. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Ruela-De-Sousa, R.R.; Fuhler, G.M.; Blom, N.; Ferreira, C.V.; Aoyama, H.; Peppelenbosch, M.P. Cytotoxicity of apigenin on leukemia cell lines: Implications for prevention and therapy. Cell Death Dis. 2010, 1, e19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Fatima, S.; Muzammal, M.; Khan, M.A.; Farid, A.; Kamran, M.; Qayum, J.; Qureshi, M.; Khan, M.N.; Khan, M.A. CRISPR/Cas9 endonucleases: A new era of genetic engineering. Abasyn J. life Sci. 2021, 29–39. [Google Scholar] [CrossRef]
  119. Seo, H.S.; Choi, H.S.; Kim, S.R.; Choi, Y.K.; Woo, S.M.; Shin, I.; Woo, J.K.; Park, S.Y.; Shin, Y.C.; Ko, S.K. Apigenin induces apoptosis via extrinsic pathway, inducing p53 and inhibiting STAT3 and NFκB signaling in HER2-overexpressing breast cancer cells. Mol. Cell. Biochem. 2012, 366, 319–334. [Google Scholar] [CrossRef]
  120. Shukla, S.; Fu, P.; Gupta, S. Apigenin induces apoptosis by targeting inhibitor of apoptosis proteins and Ku70-Bax interaction in prostate cancer. Apoptosis 2014, 19, 883–894. [Google Scholar] [CrossRef]
  121. Shukla, S.; Bhaskaran, N.; Babcook, M.A.; Fu, P.; MacLennan, G.T.; Gupta, S. Apigenin inhibits prostate cancer progression in TRAMP mice via targeting PI3K/Akt/FoxO pathway. Carcinogenesis 2014, 35, 452–460. [Google Scholar] [CrossRef] [Green Version]
  122. Karin, M. Nuclear factor-κB in cancer development and progression. Nature 2006, 441, 431–436. [Google Scholar] [CrossRef] [PubMed]
  123. Chaturvedi, M.M.; Sung, B.; Yadav, V.R.; Kannappan, R.; Aggarwal, B.B. NF-κB addiction and its role in cancer: One size does not fit all. Oncogene 2011, 30, 1615–1630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Shukla, S.; MacLennan, G.T.; Fu, P.; Patel, J.; Marengo, S.R.; Resnick, M.I.; Gupta, S. Nuclear factor-κB/p65 (Rel A) is constitutively activated in human prostate adenocarcinoma and correlates with disease progression. Neoplasia 2004, 6, 390–400. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Shukla, S.; MacLennan, G.T.; Marengo, S.R.; Resnick, M.I.; Gupta, S. Constitutive activation of PI3K-Akt and NF-κB during prostate cancer progression in autochthonous transgenic mouse model. Prostate 2005, 64, 224–239. [Google Scholar] [CrossRef] [PubMed]
  126. Lessard, L.; Karakiewicz, P.I.; Bellon-Gagnon, P.; Alam-Fahmy, M.; Ismail, H.A.; Mes-Masson, A.M.; Saad, F. Nuclear localization of nuclear factor-κB p65 in primary prostate tumors is highly predictive of pelvic lymph node metastases. Clin. Cancer Res. 2006, 12, 5741–5745. [Google Scholar] [CrossRef] [Green Version]
  127. Domingo-Domenech, J.; Mellado, B.; Ferrer, B.; Truan, D.; Codony-Servat, J.; Sauleda, S.; Alcover, J.; Campo, E.; Gascon, P.; Rovira, A.; et al. Activation of nuclear factor-κB in human prostate carcinogenesis and association to biochemical relapse. Br. J. Cancer 2005, 93, 1285–1294. [Google Scholar] [CrossRef] [Green Version]
  128. Ross, J.S.; Kallakury, B.V.S.; Sheehan, C.E.; Fisher, H.A.G.; Kaufman, R.P.; Kaur, P.; Gray, K.; Stringer, B. Expression of Nuclear Factor-κB and IκBα Proteins in Prostatic Adenocarcinomas: Correlation of Nuclear Factor-κB Immunoreactivity with Disease Recurrence. Clin. Cancer Res. 2004, 10, 2466–2472. [Google Scholar] [CrossRef] [Green Version]
  129. Shukla, S.; Kanwal, R.; Shankar, E.; Datt, M.; Chance, M.R.; Fu, P.; MacLennan, G.T.; Gupta, S. Apigenin blocks IKKa activation and suppresses prostate cancer progression. Oncotarget 2015, 6, 31216–31232. [Google Scholar] [CrossRef]
  130. Shukla, S.; Shankar, E.; Fu, P.; MacLennan, G.T.; Gupta, S. Suppression of NF-κB and NF-κB-regulated gene expression by apigenin through IκBα and IKK pathway in TRAMP mice. PLoS ONE 2015, 10, e0138710. [Google Scholar] [CrossRef] [Green Version]
  131. Masuelli, L.; Benvenuto, M.; Mattera, R.; Di Stefano, E.; Zago, E.; Taffera, G.; Tresoldi, I.; Giganti, M.G.; Frajese, G.V.; Berardi, G.; et al. In vitro and in vivo anti-tumoral effects of the flavonoid apigenin in malignant mesothelioma. Front. Pharmacol. 2017, 8, 373. [Google Scholar] [CrossRef] [Green Version]
  132. Chen, M.; Wang, X.; Zha, D.; Cai, F.; Zhang, W.; He, Y.; Huang, Q.; Zhuang, H.; Hua, Z.C. Apigenin potentiates TRAIL therapy of non-small cell lung cancer via upregulating DR4/DR5 expression in a p53-dependent manner. Sci. Rep. 2016, 6, 35468. [Google Scholar] [CrossRef] [Green Version]
  133. Krens, S.F.G.; Spaink, H.P.; Snaar-Jagalska, B.E. Functions of the MAPK family in vertebrate-development. FEBS Lett. 2006, 580, 4984–4990. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Hasnat, M.A.; Pervin, M.; Lim, J.H.; Lim, B.O. Apigenin attenuates melanoma cell migration by inducing anoikis through integrin and focal adhesion kinase inhibition. Molecules 2015, 20, 21157–21166. [Google Scholar] [CrossRef] [PubMed]
  135. Roux, P.P.; Blenis, J. ERK and p38 MAPK-Activated Protein Kinases: A Family of Protein Kinases with Diverse Biological Functions. Microbiol. Mol. Biol. Rev. 2004, 68, 320–344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Van Dross, R.T.; Hong, X.; Pelling, J.C. Inhibition of TPA-induced cyclooxygenase-2 (COX-2) expression by apigenin through downregulation of Akt signal transduction in human keratinocytes. Mol. Carcinog. 2005, 44, 83–91. [Google Scholar] [CrossRef] [PubMed]
  137. Cai, Q.; Dozmorov, M.; Oh, Y. IGFBP-3/IGFBP-3 Receptor System as an Anti-Tumor and Anti-Metastatic Signaling in Cancer. Cells 2020, 9, 1261. [Google Scholar] [CrossRef]
  138. Wang, H.; Guo, B.; Lin, S.; Chang, P.; Tao, K. Apigenin inhibits growth and migration of fibroblasts by suppressing FAK signaling. Aging 2019, 11, 3668–3678. [Google Scholar] [CrossRef]
  139. Zhao, G.; Han, X.; Cheng, W.; Ni, J.; Zhang, Y.; Lin, J.; Song, Z. Apigenin inhibits proliferation and invasion, and induces apoptosis and cell cycle arrest in human melanoma cells. Oncol. Rep. 2017, 37, 2277–2285. [Google Scholar] [CrossRef] [Green Version]
  140. Javed, Z.; Sadia, H.; Iqbal, M.J.; Shamas, S.; Malik, K.; Ahmed, R.; Raza, S.; Butnariu, M.; Cruz-Martins, N.; Sharifi-Rad, J. Apigenin role as cell-signaling pathways modulator: Implications in cancer prevention and treatment. Cancer Cell Int. 2021, 21, 189. [Google Scholar] [CrossRef]
  141. Shukla, S.; Gupta, S. Apigenin-induced cell cycle arrest is mediated by modulation of MAPK, PI3K-Akt, and loss of cyclin D1 associated retinoblastoma dephosphorylation in human prostate cancer cells. Cell Cycle 2007, 6, 1102–1114. [Google Scholar] [CrossRef]
  142. Shao, H.; Jing, K.; Mahmoud, E.; Huang, H.; Fang, X.; Yu, C. Apigenin sensitizes colon cancer cells to antitumor activity of abt-263. Mol. Cancer Ther. 2013, 12, 2640–2650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Pham, H.; Chen, M.; Takahashi, H.; King, J.; Reber, H.A.; Hines, O.J.; Pandol, S.; Eibl, G. Apigenin inhibits NNK-Induced focal adhesion kinase activation in pancreatic cancer cells. Pancreas 2012, 41, 1306–1315. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Detailed mechanism of the pharmacological properties of apigenin. (A) Anti-inflammatory response; apigenin inhibits the activity of reactive oxygen species and cancer cell metastasis which further inhibits several cell signaling pathways. It promotes the growth and development of healthy cells and downregulates the expression of cancer cell metastasis. Moreover, it blocks the activity of the phospholipase A2 (PLA2) enzyme which releases from the plasma membrane. Generally, PLA2 induces arachidonic acid (AA), which is a precursor of eicosanoid. Apigenin inhibits the expression of AA and promotes anti-inflammatory response in the host body. (B) Anti-oxidant response; host body modulates the expression of ROS when exposed to external stimuli (UV, drugs and smoke, etc.) which stimulates P53 activity and promotes apoptosis and lipid peroxidation in the host (mice) body. Apigenin consumption can significantly improve the oxidative damage caused by enhanced apoptosis and lipid peroxidation. (C) Anti-bacterial response; apigenin induces ROS in the bacterial cell and causes DNA damage and disruption in its cell wall. Furthermore, it can block beta-barrel proteins, intracellular proteins and toxins which are released from the bacterial cell. (D) Anti-viral response; apigenin enhances the reduced expression of miR122 and blocks the viral entry sites, which ultimately prevents viral DNA translation and protein synthesis. (E) Hypoglycemic response; apigenin stimulates Ras and MEK activity and increased the production of glycogen in the muscle and liver of the host body.
Figure 1. Detailed mechanism of the pharmacological properties of apigenin. (A) Anti-inflammatory response; apigenin inhibits the activity of reactive oxygen species and cancer cell metastasis which further inhibits several cell signaling pathways. It promotes the growth and development of healthy cells and downregulates the expression of cancer cell metastasis. Moreover, it blocks the activity of the phospholipase A2 (PLA2) enzyme which releases from the plasma membrane. Generally, PLA2 induces arachidonic acid (AA), which is a precursor of eicosanoid. Apigenin inhibits the expression of AA and promotes anti-inflammatory response in the host body. (B) Anti-oxidant response; host body modulates the expression of ROS when exposed to external stimuli (UV, drugs and smoke, etc.) which stimulates P53 activity and promotes apoptosis and lipid peroxidation in the host (mice) body. Apigenin consumption can significantly improve the oxidative damage caused by enhanced apoptosis and lipid peroxidation. (C) Anti-bacterial response; apigenin induces ROS in the bacterial cell and causes DNA damage and disruption in its cell wall. Furthermore, it can block beta-barrel proteins, intracellular proteins and toxins which are released from the bacterial cell. (D) Anti-viral response; apigenin enhances the reduced expression of miR122 and blocks the viral entry sites, which ultimately prevents viral DNA translation and protein synthesis. (E) Hypoglycemic response; apigenin stimulates Ras and MEK activity and increased the production of glycogen in the muscle and liver of the host body.
Molecules 27 04304 g001
Figure 2. Modulation of PI3K/AKT/MTOR pathway by apigein.
Figure 2. Modulation of PI3K/AKT/MTOR pathway by apigein.
Molecules 27 04304 g002
Figure 3. Inhibition of JAK/STAT pathway by apigenin.
Figure 3. Inhibition of JAK/STAT pathway by apigenin.
Molecules 27 04304 g003
Figure 4. Downregulation of NF-kB pathway by apigenin.
Figure 4. Downregulation of NF-kB pathway by apigenin.
Molecules 27 04304 g004
Figure 5. Inhibition of ERK/MAPK pathway by apigenin.
Figure 5. Inhibition of ERK/MAPK pathway by apigenin.
Molecules 27 04304 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Abid, R.; Ghazanfar, S.; Farid, A.; Sulaman, S.M.; Idrees, M.; Amen, R.A.; Muzammal, M.; Shahzad, M.K.; Mohamed, M.O.; Khaled, A.A.; et al. Pharmacological Properties of 4′, 5, 7-Trihydroxyflavone (Apigenin) and Its Impact on Cell Signaling Pathways. Molecules 2022, 27, 4304. https://doi.org/10.3390/molecules27134304

AMA Style

Abid R, Ghazanfar S, Farid A, Sulaman SM, Idrees M, Amen RA, Muzammal M, Shahzad MK, Mohamed MO, Khaled AA, et al. Pharmacological Properties of 4′, 5, 7-Trihydroxyflavone (Apigenin) and Its Impact on Cell Signaling Pathways. Molecules. 2022; 27(13):4304. https://doi.org/10.3390/molecules27134304

Chicago/Turabian Style

Abid, Rameesha, Shakira Ghazanfar, Arshad Farid, Samra Muhammad Sulaman, Maryam Idrees, Radwa Abdallnasser Amen, Muhammad Muzammal, Muhammad Khurram Shahzad, Mohamed Omar Mohamed, Alaa Ashraf Khaled, and et al. 2022. "Pharmacological Properties of 4′, 5, 7-Trihydroxyflavone (Apigenin) and Its Impact on Cell Signaling Pathways" Molecules 27, no. 13: 4304. https://doi.org/10.3390/molecules27134304

Article Metrics

Back to TopTop