Next Article in Journal
Gene Expression and DNA Methylation in Human Papillomavirus Positive and Negative Head and Neck Squamous Cell Carcinomas
Next Article in Special Issue
Synthetic Pathways and the Therapeutic Potential of Quercetin and Curcumin
Previous Article in Journal
New Insights into the Thermal Stability of 1-Butyl-3-methylimidazolium-Based Ionic Liquids
Previous Article in Special Issue
A Novel Drug Modulator Diarylheptanoid (trans-1,7-Diphenyl-5-hydroxy-1-heptene) from Curcuma comosa Rhizomes for P-glycoprotein Function and Apoptosis Induction in K652/ADR Leukemic Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Flavones: Six Selected Flavones and Their Related Signaling Pathways That Induce Apoptosis in Cancer

1
Research Institute of Life Science and College of Veterinary Medicine, Gyeongsang National University, 501 Jinju-daero, Jinju 52828, Korea
2
Biological Resources Research Group, Gyeongnam Department of Environment Toxicology and Chemistry, Korea Institute of Toxicology, 17 Jegok-gil, Jinju 52834, Korea
3
Division of Animal Bioscience and Intergrated Biotechnology, Jinju 52725, Korea
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2022, 23(18), 10965; https://doi.org/10.3390/ijms231810965
Submission received: 31 August 2022 / Revised: 14 September 2022 / Accepted: 15 September 2022 / Published: 19 September 2022
(This article belongs to the Special Issue Therapeutic Effects of Plant Based Bioactive Compounds on Cancer 2.0)

Abstract

:
Cancer is a horrific disease that, to date, has no cure. It is caused by various factors and takes many lives. Apoptosis is a programmed cell death mechanism and if it does not function correctly in cancer cells, it can lead to severe disease. There are various signaling pathways for regulating apoptosis in cancer cells. Flavonoids are non-artificial natural bioactive compounds that are gaining attention as being capable of for inducing apoptosis in cancer cells. Among these, in this study, we focus on flavones. Flavones are a subclass of the numerous available flavonoids and possess several bioactive functions. Some of the most reported and well-known critical flavones, namely apigenin, acacetin, baicalein, luteolin, tangeretin, and wogonin, are discussed in depth in this review. Our main aim is to investigate the effects of the selected flavones on apoptosis and cell signaling pathways that contribute to death due to various types of cancers.

1. Introduction

1.1. Cancer and Apoptosis

Cancer is one of the world’s most deadly diseases, is expensive and complicated to treat, and has many side effects due to the chemical drugs used in its treatments. The causes of cancer are diverse and can be divided into two groups: intrinsic and extrinsic risk factors. Intrinsic risks arise from accidental errors in DNA replication and cannot be modified. Extrinsic risks include both endogenous and exogenous factors. Endogenous risk factors that can be partially modified include aging, hormones, growth factors, inflammation, and DNA repair mechanisms. Exogenous risk factors are adjustable and include radiation, chemical carcinogens, tumor-causing viruses, and smoking [1].
The hallmarks of cancer cells include the continuous maintenance of proliferation signals, avoidance of growth inhibitors, activation of invasion and metastasis, induction of angiogenesis, and the possibility of permanent proliferation. In addition, the abnormality of cellular energy metabolism, avoidance of immune destruction, tumor-promoting inflammation, genomic instability, and mutation have been suggested [2]. The malfunction of apoptosis is also a cause of cancer so cancer cells with apoptotic problems do not follow this mechanism [3].
Apoptosis, along with necrosis and autophagy, is a form of programmed cell death in cancers [4]. During the apoptosis mechanisms, there are significant morphological changes such as cell shrinkage, condensation of nuclear chromatin, DNA fragmentation, blebbing, and the formation of an apoptotic body [5,6]. Subsequently, phosphatidylserine released from the membranes causes phagocytosis and death from macrophages without an inflammatory response [7,8]. Characteristic highly regulated and programmed cell death without inducing the inflammation of apoptosis plays a crucial role in cancer prevention [9,10]. There are some crucial proteins for regulating apoptosis in cancer. First, cytochrome c (Cyt c) exists in the inner mitochondrial membrane. In the presence of dATP, Cyt c is released into the cytoplasm and combines with apoptotic protease activation factor-1 (APAF-1) and pro-caspase-9 to form apoptosome, causing cell death such as apoptosis [11,12]. The Bcl-2 family comprises members that promote, inhibit, and regulate apoptosis by controlling mitochondrial outer membrane permeability [11]. Bcl-2 families divide into the anti-apoptosis members Bcl-2, Bcl-xL, Bcl-w, and Mcl-1 and pro-apoptosis members BAX, BAK, and BID [13]. The tumor-suppressor protein p53 plays an essential role in the tumorigenesis and development of tumors [14]. When Fas combines with its ligand Fas L, it activates a lethal signal then ultimately causes a series of characteristic cellular changes and leads to apoptosis. During apoptosis in specific cells, the expression of Fas increases [15]. Survivin is another apoptosis-associated protein with a potent inhibitory role in apoptosis [16]. In addition, the caspase family of proteins, namely caspase-2, -8, -9, and -10, are upstream in the portion of the cascade and are the representative initiators of apoptosis. Caspase-3, -6, and -7 are the main factors that induce apoptosis and are downstream in the portion of the apoptotic cascade. In particular, caspase-3 is the main executioner in the apoptotic process [17]. Nuclear factor-kappa B (NF-кB) is a dimeric transcription factor protein complex that regulates cell inflammation, division, and apoptosis. The IкB kinase complex (IKK) is an upstream part of the NF-кB signaling cascade, which is one of the crucial regulators of cell survival, immune response, inflammation, and tumorigenesis [18]. Serine/threonine kinase protein kinase B, also called AKT, crucially regulates the balance between cell survival and death. The PI3K/AKT pathway is an intracellular signaling pathway that promotes metabolism, proliferation, cell survival, growth, and angiogenesis in response to extracellular signals. In addition, the PI3K/AKT pathway may regulate cell cycle progression and apoptosis [19,20].
In the course of cancer treatment, many studies have been conducted, and a sufficient understanding of apoptosis is considered to be important in dealing with cancer [3].

1.2. An Overview of Flavonoids and Their Structure–Activity Relationship in Anticancer Effects

To effectively prevent cancers through apoptosis and treat them without side effects, flavonoids as phenolic compounds play a significant role as natural biological response modifiers (BRM) [21]. Flavonoids are the secondary metabolites made by plants for self-protection from external factors and are the most common compounds present in plants and fungi such as vegetables and fruits. As shown in Figure 1, flavonoids are based on a 15-carbon skeleton consisting of two benzene rings (A and B rings), and then depending on the linkage through the heterocyclic pyran ring, flavonoids can be divided into six major subclasses, which include anthocyanidins, flavones, flavan-3-ols, flavanones, flavonols, and isoflavones [22,23,24]. Currently, more than 6000 flavonoids have been identified among the subclasses. There are many studies that show flavonoids induce apoptosis through various pathways and have anticancer effects [25,26]. Many researchers anticipate that these flavonoids will bring enormous benefits to the treatment of cancer [27].
As an anticancer effect, the structure–activity relationship revealed in the comparison of flavonoids showed that the adjacent di-OH at the 3′ and 4′ of the B ring and the C2-C3 double bonds were important in the potent antiproliferative activity [28,29]. In polymethoxylated flavones, C3-OH and C8 methoxyl groups play an important role in antiproliferation [30]. Compared to flavones, the hydroxyl group of C3 of flavonols does not show antiproliferative effects, and the high activity of myricetin compared to quercetin is explained by the presence of the third OH in the B ring adjacent to the catechol group [31,32]. In other studies, the di-OH 3′ and 4′, C2-C3 double bonds, and the carbonyl group C4 were shown to play important roles in the inhibition of proliferation through cell cycle arrest [33,34]. In addition, the structure–activity relationship of anticancer action through angiogenesis inhibition shows that flavanones are more effective at inhibiting the angiogenesis factor (VEGF) than flavones or flavanols. This is because the increase in the hydroxyl group of the B ring increases the inhibitory effect. Additionally, the absence of a double bond between C2 and C3 and the level of glycosylation may play an important role in angiogenesis inhibition [31,35,36].

1.3. An Overview of Flavones and Their Biological and Pharmacological Effects

Flavones are a subclass of flavonoids (Figure 1B) and are widely distributed in the plant kingdom including in parsley, celery, red pepper, and various kinds of herbs. The molecular formula of flavones is C15H10O2 and the chemical structure is C6-C3-C6. The basic flavonoid structure contains a ketone group at C4 and has a double bond between C2 and C3 [37,38]. Apigenin, acacetin, baicalin, luteolin, tangeretin, and wogonin are formed from the flavone core [39]. Researchers have found that flavones play a role in numerous biological activities such as protecting cell membranes [40,41], and can functions as antioxidants [42,43,44], xanthine oxidase inhibitors [45], and lipid-lowering agents [46]. Antiatherogenic agents [47], nitric oxide synthase inhibitors [48], cyclo-oxygenase inhibitors [49,50], leukotriene inhibitors [51], spasmolytic agents [52], NAD(P)H quinone acceptor oxidoreductase inhibitors [53], GABA antagonists [54,55], prostate hyperplasia therapeutics [56,57], antihyperglycemic [58], α-amylase inhibitors [59,60], glycogen phosphorylase inhibitors [61], aldose reductase inhibitors [62], sirtuin activators [63], hepatoprotective agents [41,64], antiarrhythmic agents [65,66], and anti-ulcer agents [67] are some biological functions of flavones. Flavones also function as antivirals [68], antibacterials [69], antifungals [70], antiprotozoals [71], photo-protectants [72], cosmetic agents [73], and phosphodiesterase inhibitors [74] (Table 1).

1.4. Flavones as Excellent Anticancer Agents

Flavones play a crucial role as anticancer agents [23,41] (Table 1). This is the reason that we focused on flavones and their anticancer relationship through apoptosis. In addition to the above six flavones, numerous flavones have biological and pharmacological effects on cancer through mitosis inhibition and angiogenesis inhibition. An abnormal cell cycle causes excessive cell proliferation, leading to cancer, and generating limitless new blood vessels is an advantage for cancer cell growth [41]. Another example of the biological and pharmacological effects of flavones on cancer is the inhibition of the protein tyrosine kinase. The binding of various growth factors to their receptors is related to cancer development. The abnormal activation of protein tyrosine kinase involved in this binding is a major target for anticancer treatments. Therefore, the inhibition of protein tyrosine kinase could be a useful anticancer strategy. For example, luteolin blocks the ATP binding site on the catalytic unit of protein kinase C (PKC). Chrysin also inhibits tyrosine kinase activity [41]. Flavone-8-acetic acid is involved in ornithine decarboxylase inhibition and the related anticancer activity. Rutin and luteolin are involved in aromatase inhibition related to breast cancer. Combretastatin acts as a powerful tubulin polymerization inhibitor, resulting in anticancer activity (Table 1) [41].
Overall, six flavones with anticancer effects were selected from numerous flavonoids, and the anticancer effects through the signaling pathways inducing apoptosis were investigated. In this paper, we review the apoptosis and signaling pathways that form the anticancer effects of flavones. This paper could therefore be useful for researchers studying the relationship between the anticancer effects of flavones and their signaling pathways that induce apoptosis.

2. Six Selected Flavones and Their Anticancer Effects

The anticancer effects of flavones have been studied by many researchers. Flavones showing typical anticancer effects include apigenin, acacetin, baicalein, luteolin, tangeretin, and wogonin. Figure 2 shows the chemical structures of flavones. These flavones are effective in treating breast cancer, lung cancer, stomach cancer, liver cancer, skin cancer, ovarian cancer, cervical cancer, and prostate cancer, all of which have high incidences and high mortality rates worldwide [23,75,76,77,78,79].

2.1. Apigenin

Apigenin is called 4′,5,7-trihydroxyflavone (Figure 2A). The molecular formula of apigenin is C15H10O5 and its molecular weight is 270.24 g/mol [80]. It is found in many plants, fruits, and beverages such as parsley, grapes, apples, chamomile tea, and red wine [80,81]. Many studies have revealed the anticancer effects of apigenin [82]. In particular, it is a flavone that is effective against breast, prostate, liver, skin, colorectal, and lung cancers [83].

2.2. Acacetin

Acacetin, also called 5,7-dihydroxy-4′-methoxyflavone (Figure 2B), has a molecular formula of C16H12O5 and a molecular weight of 284.26 g/mol [84]. Acacetin is mainly found in the safflower, propolis, and Asteraceae plants [85]. It is effective for treating breast, stomach, gastric, and prostate cancers [78,84,85,86,87].

2.3. Baicalein

Baicalein is mainly extracted from a plant called Scutellaria baicalensis and it is called 5,6,7-trihydroxyflavone [88,89]. Baicalein has a molecular formula of C15H10O5 and a molecular weight of 270.24 g/mol [89] (Figure 2C). It is effective against breast, liver, gastric, stomach, and ovarian cancers [88,89].

2.4. Luteolin

Luteolin, highly effective for a diverse range of cancers, is also called 3,4,5,7-tetrahydroxy flavone (Figure 2D). Its molecular formula is C15H10O6 and its molecular weight is 286.24 g/mol [90]. Luteolin is mainly found in abundance in broccoli, carrots, celery, cabbage, and parsley, and it is effective against breast, lung, stomach, liver, and cervical cancers [91,92,93].

2.5. Tangeretin

Tangeretin is called 4′,5,6,7,8-penta methoxyflavone (Figure 2E) and is mainly found in the citrus family. The molecular formula of this flavone is C20H20O7 and the molecular weight is 372.37 g/mol [94]. It is effective in treating breast, gastric, prostate, and bladder cancers [95,96,97].

2.6. Wogonin

Wogonin is called 5,7-dihydroxy-8-methoxyflavone (Figure 2F). The molecular formula of this flavone is C16H12O5 and its molecular weight is 284.26 g/mol [98,99]. Wogonin is mainly extracted from Scutellaria baicalensis Georgi (Lamiaceae) and is effective in treating breast, colorectal, lung, and ovarian cancers and glioma [98,100,101].
Table 2 shows the representative flavones that are effective against various cancers.

3. Apoptosis Pathways

In cancer cells, apoptosis is a very complex mechanism that is indispensable. If apoptosis does not function properly, cancer cells proliferate indefinitely [3]. As shown in Figure 3, there are two main pathways in apoptosis; one is the intrinsic pathway and the other is the extrinsic pathway [10].

3.1. Intrinsic Pathway

In the intrinsic pathway, cytochrome c is released from the mitochondria by the Bcl-2 family, which combines with apoptotic protease activation factor-1 (APAF-1) and ATP to form an apoptosome complex by binding pro-caspase-9. It activates caspase-9 and caspase-3, -6, and -7 which are the executioners of apoptosis [10,11,102].

3.2. Extrinsic Pathway

The extrinsic pathway is mediated by the death receptors including the Fas receptor, tumor necrosis factor receptor (TNF), and TNF-related apoptosis-inducing ligand receptor (TRAIL) [103]. These death receptors have death domains called the TNF receptor-associated death domain (TRADD) and the Fas-associated death domain (FADD) [3]. Death ligands (Fas L, TNF) bind to the death receptor to form a death-inducing signaling complex (DISC), which activates caspase-8 and caspase-3, the executioner of apoptosis [3,10,103,104].
Numerous signaling pathways are involved in these two major (extrinsic and intrinsic) apoptosis pathways. The regulation of apoptosis is imperative in cancer cells as well as some of the major cell signaling pathways involved in the survival and death of cancer cells.

4. Signaling Pathways Related to Cancer Cell Apoptosis

4.1. PI3K/AKT Pathway

The PI3K/AKT pathway, which can regulate apoptosis in cancer cells, plays a major role in mammalian cell proliferation, differentiation, autophagy, survival, and apoptosis [105]. The lipid kinase phosphoinositide 3-kinase (PI3K) consists of p110, a catalytic unit, and p85, a regulatory unit, which phosphorylates inositol carbon 3 of phosphatidylinositol 4,5-bisphosphate (PIP2) to phosphatidylinositol 3,4,5-triphosphate (PIP3) [106,107]. When PIP3 is increased, AKT and phosphoinositide-dependent kinase 1 (PDK1), which are enzymes in the pleckstrin homology domain (PH), are gathered and AKT is activated by PDK1 and mTOR complex 2 (mTORC2) [106,107]. This activated AKT phosphorylates BAD, caspase-9, and FOXO. Then, these proteins are inactivated, leading to the inhibition of apoptosis [108,109,110]. The activation and inhibition of the PI3K/AKT and mTOR pathways in human cancer cells determine the survival, carcinogenicity, metastasis, and invasion of cancer cells [111,112]. In cancer cells, this pathway is highly activated, and the inhibition of the PI3K/AKT and mTOR pathways induces apoptosis and has anticancer effects in various cancers [113].

4.2. Wnt Pathway

The Wnt pathway is a signaling pathway involved in cell proliferation, migration, stem cell differentiation, and various diseases including cancer. In this pathway, β-catenin belongs downstream of Wnt and plays a key role in apoptosis while triggering transcription by the Wnt ligand [114,115,116]. Wnt acts as a ligand in this pathway and its receptor is Frizzled [114]. In the absence of Wnt, destruction complexes such as Axin, glycogen synthase kinase-3 beta (GSK3β), adenomatous polyposis coli protein (APC), and casein kinase (CK) phosphorylate β-catenin. Due to proteasome protein degradation, β-catenin is degraded in the cytoplasm and cannot meet the T-cell factor (TCF) so transcription does not occur [114,117]. On the other hand, when the Wnt ligand binds to its receptor Frizzled, phosphorylation of the lipoprotein receptor-related protein (LRP), a co-receptor of Wnt, occurs, and Axin is phosphorylated and moves to the cell together with CK and GSK3β [114,115]. The β-catenin, freed from the influence of GSK3β, enters the nucleus and meets TCF to initiate β-catenin-dependent gene expression [114,117]. Inhibition of the Wnt/β-catenin signaling pathway showed anticancer effects by attenuating survival signals and inducing apoptosis [117,118]. Another mechanism is that β-catenin downregulates anti-apoptosis proteins and induces apoptosis by including transcription of p53 and c-myc, thus inducing the transcription of pro-apoptosis proteins [119].

4.3. JAK/STAT Pathway

Abnormal activation of the janus kinase (JAK)/signal transducer and activators of the transcription (STAT) pathway promotes tumorigenesis [120]. The cytokine receptors do not have tyrosine kinase activity so there is a problem in signal transduction. When cytokines, such as chemokines, interleukins, interferon, and TNF-α, a protein immunomodulatory agent secreted from immune cells, bind to a cytokine receptor without tyrosine kinase activity, JAK is gathered around the cytokine receptor to transmit a signal [121,122,123]. In this way, JAK phosphorylates STAT to form a STAT dimer and enters the nucleus to express the STAT target gene [123]. Blocking this JAK/STAT pathway in cancer cells suppresses the expression of the target gene that controls essential cellular function and prevents the avoidance of apoptosis, a characteristic of cancer cells, and an anticancer effect can be expected [124].

4.4. MAPK Pathway

The mitogen-activated protein kinase (MAPK) pathway plays an important role in cell proliferation, differentiation, angiogenesis, tumor metastasis, and apoptosis [125]. MAP kinase kinase kinase (MAPKKK) phosphorylates MAP kinase kinase (MAPKK) by external growth factors, stress, and cytokines. MAPKK phosphorylates MAPK, and MAPK phosphorylates and activates the target protein [125,126]. MAPKKK, MAPKK, and MAPK can be divided into four major signaling systems. The four-member classes in MAPK are extracellular signaling-regulated kinase1/2 (ERK1/2), ERK5, c-Jun NH-2-terminal kinase (JNK), and p38-MAPK. There are ERK1/2, ERK5, JNK1/2/3, and p38-MAPKα/β/γ/δ sub-families in each family [125,126,127]. Among them, ERK is known to have anti-apoptotic action, and JNK and p38-MAPK are likely to act in both anti- and pro-apoptosis [127]. An increase in ERK1/2/5 in cancer cells has been confirmed in the MAPK pathway, which could be a major target for anticancer effects [127,128,129]. The activation of JNK and p38-MAPK normally promotes but does not necessarily lead to apoptosis, and both JNK and p38-MAPK mediate anti-apoptosis and pro-apoptosis [130]. To explain apoptosis in the MAPK pathway, it can be anti- or pro-apoptosis depending on the response of the target protein to various stimuli or the molecular context according to the cell type [130].

4.5. p53

As a tumor suppressor, p53 is activated when DNA is damaged, causing cell cycle arrest and apoptosis. When this protein is mutated, cells that should die do not die and cancer occurs [131]. In the normal apoptosis process in the p53 pathway, when cellular stress signals occur, p53 induces the pro-apoptotic member BH-3-only proteins (BIM, NOXA, PUMA, and BAX) [132,133,134,135]. After this, BIM, NOXA, and PUMA suppress the Bcl-2 family of anti-apoptotic proteins, and apoptosis occurs through an intrinsic pathway by activated BAX and BAK [132,133,134].
Figure 4 shows the signaling pathways that are involved in apoptosis, which are interconnected. Numerous signaling pathways are involved in this apoptosis mechanism.

5. Regulation of Signaling Pathways in Which Six Selected Flavones Induce Apoptosis in Cancer Cells

5.1. Apigenin

Apigenin has exhibited anticancer effects on many cancer cells such as breast, prostate, liver, colorectal, skin, and lung cancers [83]. Apigenin induced apoptosis in cancer cells by downregulating various signaling pathways such as the PI3K/AKT pathway, ERK1/2, NF-кB, JAK/STAT, and Wnt/β-catenin [81]. First, in breast cancer, there was no induction of apoptosis through the intrinsic pathway because the mitochondrial membrane potential was not reduced in human epidermal growth factor receptor 2(HER-2)-expressing breast cancer BT-474 cells without the effect of Bcl-2 or BAX. However, there was an apoptotic process leading to the cleavage of caspase-8 and PARP via an extrinsic pathway. In addition, it has been demonstrated that breast cancer cells induce apoptosis through the inhibition of STAT3 signaling [136]. Apigenin also revealed that HER-2-expressing MCF-7 cells, another type of breast cancer cell, were involved in an apoptosis process that induced caspase-8 and PARP cleavage and apoptosis through a p53-dependent pathway [137]. In the case of prostate cancer, apigenin decreased Bcl-2 and Bcl-xL and increased BAX in PC-3 and DU145 human prostate cancer cells. In addition, the apoptosis inhibitors XIAP, c-IAP1, c-IAP2, and survivin were inhibited, resulting in apoptosis [138]. As a result of treatment with apigenin in Hep G2 cells, which are liver cancer cells, apoptosis was induced through the inhibition of the PI3K/AKT/mTOR pathway [139]. In HCT-116 colorectal cancer cells, apigenin increased tumor-suppressor proteins such as p53 and p21 [140]. In the case of SW480 cells, apoptosis was induced by increasing caspase-3 and BAX and decreasing Bcl-2 [141,142]. In addition, apigenin induced apoptosis by inhibiting the phosphorylation of STAT3 in colon cancer cells and simultaneously downregulating anti-apoptotic proteins such as Bcl-xL and Mcl-1 [143]. Apigenin induced apoptosis by regulating AKT and MAPK in A375SM human melanoma cells [144]. In A549 lung cancer cells, caspase-3, -8, and -9 were induced by apigenin and promoted cytochrome c, and apoptosis-inducing factor (AIF) was promoted through the mitochondria and caspase-3 and -9 were activated [145]. TRAIL-induced apoptosis by the upregulation of death receptors 4 and 5 occurred in p53-dependent NSCLC lung cancer cells. In addition, BAD and BAX were upregulated; Bcl-xL and Bcl-2 were downregulated; and NF-кB, AKT, and ERK activation were inhibited [146].

5.2. Acacetin

Acacetin induced apoptosis in cancer cells through various pathways such as mitochondria-mediated death signaling, caspase activation, the β-catenin pathway, and NF-кB/AKT signaling for breast, gastric, colorectal, and prostate cancers [78,84,85,86,87]. In breast cancer cells, MCF-7 and acacetin decreased Bcl-2 and released cytochrome c and AIF through the loss of mitochondrial membrane potential. Stress-activated protein kinase/c-Jun NH-4 terminal kinase1/2 (SAPK/JNK1/2) and c-Jun were activated and then apoptosis was induced through the SAPK/JNK1/2-c-Jun pathway [87]. In gastric cancer AGS cells, an apoptotic pathway was formed by caspase activity through ROS generation, mitochondrial mediation, and Fas activation [78]. In SW480 and HCT-116 colorectal cancer cells, the downregulation of the β-catenin pathway and increase in the mitochondrial membrane potential depolarization increased the BAX:Bcl-2 ratio and although there was no change in caspase, apoptosis was induced by AIF [86]. In DU145 prostate cancer cells, acacetin reduced phospho-AKT in a concentration-dependent manner, and phospho-GSK-3β, downstream of AKT, decreased and p53 increased. In addition, the activity of phospho-IкB and NF-кB was decreased, and apoptosis was induced through the NF-кB/AKT pathway due to the decrease in XIAP, and Bcl-2 [84].

5.3. Baicalein

Baicalein regulated apoptosis through various pathways for breast, liver, stomach, colorectal, lung, cervical, and ovarian cancers [88,89]. Baicalein regulated mitochondrial potential in MDA-MB-231 breast cancer cells, released cytochrome c, and activated caspase-3 to induce apoptosis [147]. In addition, increased p53 expression and increased ERK/p38 MAPK in MDA-MB-231 cells were associated with the pro-apoptotic effect of baicalein [89]. Additionally, in breast cancer cells MCF-7 and MDA-MB-231, Bcl-2 decreased and BAX increased. So, it induced apoptosis and autophagy through the inhibition of the PI3K/AKT pathway and downregulated the expression of phospho-AKT, phospho-mTOR, NF-кB, and phospho-IкB [147,148]. In HCC (hepatocellular carcinoma) liver cancer cells, baicalein increased BAX, decreased Bcl-2, and induced cleaved caspase-3, -9, and PARP. In addition, baicalein induced apoptosis by activating JNK [149]. In SGC-7901 gastric cancer cells, apoptosis was induced through the mitochondrial pathway. The mitochondrial membrane potential was disrupted so Bcl-2 decreased and BAX increased [150]. In HCT-116 and SW480 colon cancer cells, baicalein was involved in apoptosis by the MAPK/ERK and p38 pathways [88]. In A2780 ovarian cancer cells, baicalein decreased Bcl-2 and activated caspase-3 and -9 [151].

5.4. Luteolin

Luteolin induced apoptosis in breast, lung, gastric, liver, and cervical cancers [91,92,93]. First, in breast cancer, luteolin inhibited PI3K/AKT activation and increased FOXO3a activation, leading to cell cycle arrest and apoptosis [152]. Additionally, luteolin induced apoptosis through the downregulation of the human telomerase reverse transcriptase (hTERT) [153]. In SCLC lung cancer cells, luteolin caused cisplatin to become more susceptible to anticancer effects through JNK-activated apoptosis [154]. In NSCLC lung cancer cells, the expression of BAX, p53, p21, caspase-3, and -9 decreased and Bcl-2 increased due to microRNA-34a-5p inhibition. So, activation of microRNA-34a-5p blocked the caspase pathway and can be considered to cause apoptosis of lung cancer [155]. In BGC-823 gastric cancer cells, caspase-3, -9, and cytochrome c increased, the ratio of BAX to Bcl-2 increased, and apoptosis was induced through the inhibition of the MAPK and PI3K signaling pathways [156]. In SMMC-7721 liver cancer cells, luteolin increased caspase-8 and decreased Bcl-2 [75], and in Hep-G2 cells, it was reported that apoptosis was induced through BAX/BAK mitochondrial translocation and JNK activation [157]. In HeLa cervical cancer cells, luteolin increased the expression of various pro-apoptotic proteins, including BAX, BAD, BID, APAF1, TRADD, FAS, FADD, and caspase-3 and -9, but the expression of anti-apoptotic proteins, such as BCL-2 and MCL-1, decreased. Additionally, apoptosis occurred through the inhibition of the AKT and MAPK pathways due to the downregulation of MAPK, ERK1/2, and AKT and the upregulation of p53 at the transcriptional level [93].

5.5. Tangeretin

Tangeretin induced apoptosis in breast, gastric, prostate, and bladder cancers through various pathways [95,96,97]. The treatment of drug-resistant MDA-MB-231 breast cancer cells with tangeretin induced apoptosis by increasing BAX, caspase-3, and -8 and decreasing Bcl-2 [158]. In AGS gastric cancer cells, caspase-3, -8, and -9, as well as BAX, tBID, and p53, were upregulated and apoptosis was induced through Fas/Fas L and the p53-dependent mitochondrial pathway [95]. In PC-3 and LNCaP prostate cancer cells, caspase-3 and cleaved caspase-3 and-9 were regulated. Only PC-3, BAX, and Bcl were regulated. Apoptosis was induced by inhibiting epithelial-mesenchymal transition (EMT), a cellular program important for cancer progression in these prostate cancer cells. In DU145 prostate cancer cells, apoptosis and DNA cleavage were induced through the regulation of the androgen receptor (AR)-PI3K/AKT/mTOR-Notch signaling pathway. In BFTC-905 bladder cancer cells, apoptosis was induced by regulating the release of cytochrome c and AIF from the mitochondria and cleaved caspase-3 and -9 and pro-caspase-3 and -9 [97].

5.6. Wogonin

Wogonin induced apoptosis through various pathways in breast cancer, colorectal cancer, lung cancer, ovarian cancer, and glioma [101]. Wogonin induced apoptosis by decreasing Bcl-2 and survivin and increasing BAX, p53, and caspase-3, -8, and -9 in MCF-7 breast cancer cells. In addition, apoptosis was induced through the inhibition of the PI3K/AKT/survivin signaling pathway and ERK activation [159]. In HT-29 colorectal cancer cells, Bcl-2 expression was reduced, BAX expression increased, and apoptosis was induced through the PI3K/AKT pathway [160]. In addition, in HCT116 cells, cell cycle arrest through the inhibition of the β-catenin-dependent Wnt signaling pathway showed attenuated proliferation and apoptosis in colorectal cancer [161]. In SW48, another colorectal cancer cell, apoptosis was induced by an increase in BAX and caspase-3, -8, and -9 and the inhibition of PI3K/AKT/STAT-3 [162]. In A549 lung cancer cells, wogonin induced apoptosis associated with increases in caspase-3, -8, and -9 and PARP and decreases in XIAP, c-FLIP, cIAP-1, -2, AKT, and STAT-3 [163]. In A2780 ovarian cancer cells, apoptosis was induced due to an increase in BAX and p53 and a decrease in AKT and Bcl-2 [98]. In U251 and U87 glioblastoma cells, wogonin induced not only the upregulation of caspase-3 and-9 but also PARP expression. In addition, it induced the generation of reactive oxygen species (ROS) and activated endoplasmic reticulum (ER) stress to induce apoptosis [101].
Comprehensively, Figure 5 shows the cell signaling regulation of apoptosis by apigenin, acacetin, baicalein, luteolin, tangeretin, and wogonin. Table 3 also shows the cancer cell lines, treatment concentrations, and apoptotic regulation of cell signaling pathways associated with the six above selected flavones.

6. Anticancer Effects of Six Selected Flavones in an In Vivo Model

Apigenin suppressed colorectal cancer and reduced tumor volume in prostate cancer without reducing appetite when ingested in Sprague–Dawley (SD) rats. Colorectal cancer was induced by azoxymethane (AOM), which is effective for the induction of colon carcinoma in SD rats. In these SD mice, apigenin treatment had the potential to suppress colorectal cancer by reducing aberrant crypt foci (ARF), which are clusters of abnormal tube-like glands, and increasing apoptosis [141]. In addition, in transgenic adenocarcinoma of mouse prostate (TRAMP) mice, apigenin inhibited the formation of prostate carcinoma by regulating the PI3K/AKT/FOXO pathway [164,165]. However, low bioavailability and instability in the gut were present [166,167]. Amounts of 20 and 50 μg of apigenin were administered to a xenograft model in which PC-3 prostate cancer cells were implanted into athymic nude mice. This resulted in a significant decrease in protein expression of XIAP and survivin in PC-3 tumor xenografts. A dose-dependent decrease in HDAC1 expression and an increase in Bax and PARP cleavage were observed in apigenin-treated mice [138].
Acacetin inhibited tumor growth through STAT3 regulation in DU145 prostate cancer cells in a nude mouse xenograft model [168].
As a result of analyzing the gene expression of a lung cancer xenograft tumor mouse model, the genes when baicalein induce apoptosis, ITGB3 (+6.96) and TNFRSF25 (+3.4), were most significantly upregulated [77]. In another study, mTOR inhibition prevented tumor growth in a breast cancer xenograft mouse model [169]. In addition, cervical cancer cell apoptosis was induced through an increase in BAX and a decrease in Bcl-2 in an in vivo tumor model [170].
Luteolin reduced the weight and volume of gastric tumors in a rat model by the inhibition of Notch1 and β-catenin [171]. When luteolin was administered to the lung cancer xenograft tumor mouse model, tumor suppression was shown by the inhibition of mutant EGF receptors, a major oncogene that induces tumorigenesis in many types of cancer, including NSCLC. This was followed by inducing the degradation of the EGF receptors and then preventing the PI3K/AKT/mTOR signal, which resulted in apoptosis and toxicity, and weight loss did not occur [172].
Tangeretin inhibited tumor growth in the MDA-MB-231 breast cancer cell nude mouse xenograft model [94]. Tangeretin had a strong inhibitory effect on tumor growth in the SGC-7901 gastric cancer cell nude mouse xenograft model [173].
As a result of treatment with wogonin in BALB/C nude mice xenografted with A2780 ovarian cancer cells and HT-29 colorectal cancer cells, the tumor volume and weight were reduced [174].
Table 4 shows the in vivo functions of anticancer effects in various cancer types.

7. Discussion and Conclusions

7.1. The Prospects of Compounds Such as Flavones and Flavonoids for Anticancer Effects

Cancer is still a disease whose treatments face many challenges. Anticancer research on polyphenols, such as flavonoids, has been actively conducted over the past few decades [175]. Of course, among flavonoids, flavones cannot be said to have the greatest anticancer effects and their mechanisms have not been fully elucidated. However, as shown in this review, flavones are effective in treating various cancers and they have special qualities that make them more effective than other flavonoids. This provides ample evidence to support flavones as having good prospects for treating cancer in the future so they are worth studying. However, in order to more clearly understand the effects of flavones and flavonoids, further studies are needed that focus on treatment effects, diet and bioavailability, absorption, metabolism, physiologically relevant models, and structural relationships between flavones and flavonoid subclasses [176]. In addition, it is important to proceed with clinical research using these data. There have been some clinical trials involving flavone-based agents, such as flavopiridol, which inhibits cdk and induces apoptosis, but more trials are needed [177].

7.2. Noteworthy ER-Mediated Apoptosis and the Need to Study Related Flavones

In this review, the apoptosis of the external and internal pathways has been mainly investigated, but the endoplasmic reticulum (ER)-mediated apoptosis pathway is also emerging as an important apoptosis pathway. The endoplasmic reticulum is an important organelle of eukaryotic cells and is necessary for maintaining intracellular homeostasis; controlling the synthesis, secretion, and folding of membrane-bound proteins; and transporting calcium ions [178]. When a misfolded protein accumulates in the ER and leads to activation of the unfolded protein response (UPR), it is called ER stress [178,179]. Through this ER stress, the combination with the mitochondria plays an important role in regulating apoptosis [178]. The first key protein in the ER stress-mediated apoptosis pathway is the C/EBP homologous protein (CHOP). The activation of the protein kinase R-like ER kinase (PERK) phosphorylates eukaryotic initiation factor 2 (EIF2a) and the activating transcription factor 4 (ATF4) is activated to upregulate the CHOP. The expression of this CHOP is also increased by activating transcription factor 6 (ATF6). The CHOP subsequently decreases the expression of anti-apoptotic Bcl-xL and Bcl-2 and conversely increases the pro-apoptotic BAX, BAK, NOXA, BIM, and PUMA [180,181]. Another pathway is that inositol-requiring enzyme 1 (IRE1) binds to tumor necrosis factor (TNF) receptor-associated factor 2 (TRAF2) to activate caspase 12, which in turn, activates caspase-3. In addition, the IRE1α-TRAF2 complex activates apoptosis signal-regulating kinase 1 (ASK1) and induces phosphorylation of JNK to induce apoptosis through c-jun, activates BAX and BAK, and inactivates Bcl-2 to play an important role in apoptosis [182].
Therefore, this endoplasmic reticulum-mediated apoptosis is very important and a pathway in which natural compounds, such as flavones, control ER stress [178] to act as an anticancer agent, which can be an important approach in the fight against cancer. Studies have been conducted on natural compounds, such as several flavonoids, but more research is needed.

7.3. Conclusions

This review includes a basic introduction to the selected flavones and their anticancer effects. In addition, we investigated the various pathways involved in apoptosis and discussed the signaling pathways by which flavones induce apoptosis. Flavones could be promising compounds for anticancer action in the future after further research considering this review information. Lastly, we hope this review will serve as a valuable manual for anticancer-related research using flavones.

Author Contributions

S.H.J. and H.H.K. conceptualization, writing—review and editing; S.H.J. writing—original draft preparation; S.E.H. and M.Y.P. participated in the focus group discussions; P.B.B. and A.A. validation; K.I.P., J.D.H. and H.W.K. project administration; G.S.K. supervision. All authors have read and agreed to the published version of the manuscript.

Funding

The present study was supported by grants from the National Research Foundation funded by the Ministry of Science and ICT, Republic of Korea (grant no. 2020R1A2B5B01001807).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Wu, S.; Zhu, W.; Thompson, P.; Hannun, Y.A. Evaluating intrinsic and non-intrinsic cancer risk factors. Nat. Commun. 2018, 9, 3490. [Google Scholar] [CrossRef]
  2. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef]
  3. Wong, R.S. Apoptosis in cancer: From pathogenesis to treatment. J. Exp. Clin. Cancer Res. 2011, 30, 87. [Google Scholar] [CrossRef]
  4. Kroemer, G.; Galluzzi, L.; Vandenabeele, P.; Abrams, J.; Alnemri, E.S.; Baehrecke, E.H.; Blagosklonny, M.V.; El-Deiry, W.S.; Golstein, P.; Green, D.R.; et al. Classification of cell death: Recommendations of the Nomenclature Committee on Cell Death 2009. Cell Death Differ. 2009, 16, 3–11. [Google Scholar] [CrossRef]
  5. Kerr, J.F.; Wyllie, A.H.; Currie, A.R. Apoptosis: A basic biological phenomenon with wide-ranging implications in tissue kinetics. Br. J. Cancer 1972, 26, 239–257. [Google Scholar] [CrossRef]
  6. Kerr, J.F.; Winterford, C.M.; Harmon, B.V. Apoptosis. Its significance in cancer and cancer therapy. Cancer 1994, 73, 2013–2026. [Google Scholar] [CrossRef]
  7. Hacker, G. The morphology of apoptosis. Cell Tissue Res. 2000, 301, 5–17. [Google Scholar] [CrossRef]
  8. Pistritto, G.; Trisciuoglio, D.; Ceci, C.; Garufi, A.; D’Orazi, G. Apoptosis as anticancer mechanism: Function and dysfunction of its modulators and targeted therapeutic strategies. Aging 2016, 8, 603–619. [Google Scholar] [CrossRef]
  9. Haanen, C.; Vermes, I. Apoptosis and inflammation. Mediat. Inflamm. 1995, 4, 5–15. [Google Scholar] [CrossRef]
  10. Pfeffer, C.M.; Singh, A.T.K. Apoptosis: A Target for Anticancer Therapy. Int. J. Mol. Sci. 2018, 19, 448. [Google Scholar] [CrossRef] [Green Version]
  11. Kroemer, G.; Galluzzi, L.; Brenner, C. Mitochondrial membrane permeabilization in cell death. Physiol. Rev. 2007, 87, 99–163. [Google Scholar] [CrossRef]
  12. Li, P.; Nijhawan, D.; Budihardjo, I.; Srinivasula, S.M.; Ahmad, M.; Alnemri, E.S.; Wang, X. Cytochrome c and dATP-dependent formation of Apaf-1/caspase-9 complex initiates an apoptotic protease cascade. Cell 1997, 91, 479–489. [Google Scholar] [CrossRef]
  13. Levine, B.; Sinha, S.; Kroemer, G. Bcl-2 family members: Dual regulators of apoptosis and autophagy. Autophagy 2008, 4, 600–606. [Google Scholar] [CrossRef]
  14. Zhang, Z.; Wang, C.Z.; Du, G.J.; Qi, L.W.; Calway, T.; He, T.C.; Du, W.; Yuan, C.S. Genistein induces G2/M cell cycle arrest and apoptosis via ATM/p53-dependent pathway in human colon cancer cells. Int. J. Oncol. 2013, 43, 289–296. [Google Scholar] [CrossRef]
  15. Waring, P.; Mullbacher, A. Cell death induced by the Fas/Fas ligand pathway and its role in pathology. Immunol. Cell Biol. 1999, 77, 312–317. [Google Scholar] [CrossRef]
  16. Jaiswal, P.K.; Goel, A.; Mittal, R.D. Survivin: A molecular biomarker in cancer. Indian J. Med. Res. 2015, 141, 389–397. [Google Scholar]
  17. Li, J.; Yuan, J. Caspases in apoptosis and beyond. Oncogene 2008, 27, 6194–6206. [Google Scholar] [CrossRef]
  18. Solt, L.A.; May, M.J. The IkappaB kinase complex: Master regulator of NF-kappaB signaling. Immunol. Res. 2008, 42, 3–18. [Google Scholar] [CrossRef]
  19. Nicholson, K.M.; Anderson, N.G. The protein kinase B/Akt signalling pathway in human malignancy. Cell Signal. 2002, 14, 381–395. [Google Scholar] [CrossRef]
  20. Nitulescu, G.M.; Van De Venter, M.; Nitulescu, G.; Ungurianu, A.; Juzenas, P.; Peng, Q.; Olaru, O.T.; Gradinaru, D.; Tsatsakis, A.; Tsoukalas, D.; et al. The Akt pathway in oncology therapy and beyond (Review). Int. J. Oncol. 2018, 53, 2319–2331. [Google Scholar] [CrossRef]
  21. Abotaleb, M.; Samuel, S.M.; Varghese, E.; Varghese, S.; Kubatka, P.; Liskova, A.; Busselberg, D. Flavonoids in Cancer and Apoptosis. Cancers 2018, 11, 28. [Google Scholar] [CrossRef] [Green Version]
  22. Raffa, D.; Maggio, B.; Raimondi, M.V.; Plescia, F.; Daidone, G. Recent discoveries of anticancer flavonoids. Eur. J. Med. Chem. 2017, 142, 213–228. [Google Scholar] [CrossRef]
  23. Kopustinskiene, D.M.; Jakstas, V.; Savickas, A.; Bernatoniene, J. Flavonoids as Anticancer Agents. Nutrients 2020, 12, 457. [Google Scholar] [CrossRef]
  24. Kumar, S.; Pandey, A.K. Chemistry and biological activities of flavonoids: An overview. Sci. World J. 2013, 2013, 162750. [Google Scholar] [CrossRef]
  25. Panche, A.N.; Diwan, A.D.; Chandra, S.R. Flavonoids: An overview. J. Nutr. Sci. 2016, 5, e47. [Google Scholar] [CrossRef]
  26. Ramos, S. Effects of dietary flavonoids on apoptotic pathways related to cancer chemoprevention. J. Nutr. Biochem. 2007, 18, 427–442. [Google Scholar] [CrossRef]
  27. Liskova, A.; Samec, M.; Koklesova, L.; Brockmueller, A.; Zhai, K.; Abdellatif, B.; Siddiqui, M.; Biringer, K.; Kudela, E.; Pec, M.; et al. Flavonoids as an effective sensitizer for anti-cancer therapy: Insights into multi-faceted mechanisms and applicability towards individualized patient profiles. EPMA J. 2021, 12, 155–176. [Google Scholar] [CrossRef]
  28. Kawaii, S.; Tomono, Y.; Katase, E.; Ogawa, K.; Yano, M. Antiproliferative activity of flavonoids on several cancer cell lines. Biosci. Biotechnol. Biochem. 1999, 63, 896–899. [Google Scholar] [CrossRef]
  29. Sanchez, I.; Calderon, J.; Ruiz, B.; Tellez, J.; Calzada, L.; Taboada, J. In vitro cytotoxicity of flavonoids against MK2 and C6 tumour cells. Phytother. Res. 2001, 15, 290–293. [Google Scholar] [CrossRef]
  30. Jayaprakasam, B.; Vareed, S.K.; Olson, L.K.; Nair, M.G. Insulin secretion by bioactive anthocyanins and anthocyanidins present in fruits. J. Agric. Food Chem. 2005, 53, 28–31. [Google Scholar] [CrossRef]
  31. Benavente-Garcia, O.; Castillo, J. Update on uses and properties of citrus flavonoids: New findings in anticancer, cardiovascular, and anti-inflammatory activity. J. Agric. Food Chem. 2008, 56, 6185–6205. [Google Scholar] [CrossRef] [PubMed]
  32. Manthey, J.A.; Guthrie, N. Antiproliferative activities of citrus flavonoids against six human cancer cell lines. J. Agric. Food Chem. 2002, 50, 5837–5843. [Google Scholar] [CrossRef] [PubMed]
  33. Agullo, G.; Gamet-Payrastre, L.; Manenti, S.; Viala, C.; Remesy, C.; Chap, H.; Payrastre, B. Relationship between flavonoid structure and inhibition of phosphatidylinositol 3-kinase: A comparison with tyrosine kinase and protein kinase C inhibition. Biochem. Pharmacol. 1997, 53, 1649–1657. [Google Scholar] [CrossRef]
  34. Casagrande, F.; Darbon, J.M. Effects of structurally related flavonoids on cell cycle progression of human melanoma cells: Regulation of cyclin-dependent kinases CDK2 and CDK1. Biochem. Pharmacol. 2001, 61, 1205–1215. [Google Scholar] [CrossRef]
  35. Albini, A.; Tosetti, F.; Benelli, R.; Noonan, D.M. Tumor inflammatory angiogenesis and its chemoprevention. Cancer Res. 2005, 65, 10637–10641. [Google Scholar] [CrossRef]
  36. Kim, J.D.; Liu, L.; Guo, W.; Meydani, M. Chemical structure of flavonols in relation to modulation of angiogenesis and immune-endothelial cell adhesion. J. Nutr. Biochem. 2006, 17, 165–176. [Google Scholar] [CrossRef]
  37. Gürler, S.B.; Kiraz, Y.; Baran, Y. Chapter 21—Flavonoids in cancer therapy: Current and future trends. In Biodiversity and Biomedicine; Ozturk, M., Egamberdieva, D., Pešić, M., Eds.; Academic Press: Cambridge, MA, USA, 2020; pp. 403–440. [Google Scholar]
  38. Hostetler, G.L.; Ralston, R.A.; Schwartz, S.J. Flavones: Food Sources, Bioavailability, Metabolism, and Bioactivity. Adv. Nutr. 2017, 8, 423–435. [Google Scholar] [CrossRef]
  39. Martens, S.; Mithofer, A. Flavones and flavone synthases. Phytochemistry 2005, 66, 2399–2407. [Google Scholar] [CrossRef]
  40. Ollila, F.; Halling, K.; Vuorela, P.; Vuorela, H.; Slotte, J.P. Characterization of flavonoid-biomembrane interactions. Arch. Biochem. Biophys. 2002, 399, 103–108. [Google Scholar] [CrossRef]
  41. Verma, A.K.; Pratap, R. The biological potential of flavones. Nat. Prod. Rep. 2010, 27, 1571–1593. [Google Scholar] [CrossRef]
  42. Bors, W.; Heller, W.; Michel, C.; Saran, M. Flavonoids as antioxidants: Determination of radical-scavenging efficiencies. Methods Enzymol. 1990, 186, 343–355. [Google Scholar] [PubMed]
  43. Cao, G.; Sofic, E.; Prior, R.L. Antioxidant and prooxidant behavior of flavonoids: Structure-activity relationships. Free Radic. Biol. Med. 1997, 22, 749–760. [Google Scholar] [CrossRef]
  44. Su, Y.L.; Huang, Y.; Chen, Z.Y. Isolation and elucidation of antioxidant constituents from acetone extract in root of Scutellaria rehderiana. Zhongguo Zhong Yao Za Zhi 2004, 29, 863–864. [Google Scholar] [PubMed]
  45. Lin, S.; Zhang, G.; Liao, Y.; Pan, J.; Gong, D. Die.etary Flavonoids as Xanthine Oxidase Inhibitors: Structure-Affinity and Structure-Activity Relationships. J. Agric. Food Chem. 2015, 63, 7784–7794. [Google Scholar] [CrossRef] [PubMed]
  46. Zeka, K.; Ruparelia, K.; Arroo, R.R.J.; Budriesi, R.; Micucci, M. Flavonoids and Their Metabolites: Prevention in Cardiovascular Diseases and Diabetes. Diseases 2017, 5, 19. [Google Scholar] [CrossRef]
  47. Basu, A.; Das, A.S.; Majumder, M.; Mukhopadhyay, R. Antiatherogenic Roles of Dietary Flavonoids Chrysin, Quercetin, and Luteolin. J. Cardiovasc. Pharmacol. 2016, 68, 89–96. [Google Scholar] [CrossRef]
  48. Olszanecki, R.; Gebska, A.; Kozlovski, V.I.; Gryglewski, R.J. Flavonoids and nitric oxide synthase. J. Physiol. Pharmacol. 2002, 53 Pt 1, 571–584. [Google Scholar]
  49. Liang, Y.C.; Tsai, S.H.; Tsai, D.C.; Lin-Shiau, S.Y.; Lin, J.K. Suppression of inducible cyclooxygenase and nitric oxide synthase through activation of peroxisome proliferator-activated receptor-gamma by flavonoids in mouse macrophages. FEBS Lett. 2001, 496, 12–18. [Google Scholar] [CrossRef]
  50. Chi, Y.S.; Cheon, B.S.; Kim, H.P. Effect of wogonin, a plant flavone from Scutellaria radix, on the suppression of cyclooxygenase-2 and the induction of inducible nitric oxide synthase in lipopolysaccharide-treated RAW 264.7 cells. Biochem. Pharmacol. 2001, 61, 1195–1203. [Google Scholar]
  51. Zwaagstra, M.E.; Timmerman, H.; van de Stolpe, A.C.; de Kanter, F.J.; Tamura, M.; Wada, Y.; Zhang, M.Q. Synthesis and structure-activity relationships of carboxyflavones as structurally rigid CysLT1 (LTD4) receptor antagonists. J. Med. Chem. 1998, 41, 1428–1438. [Google Scholar] [CrossRef]
  52. Chapple, C.R.; Parkhouse, H.; Gardener, C.; Milroy, E.J. Double-blind, placebo-controlled, cross-over study of flavoxate in the treatment of idiopathic detrusor instability. Br. J. Urol. 1990, 66, 491–494. [Google Scholar] [CrossRef] [PubMed]
  53. Chen, S.; Deng, P.S.; Swiderek, K.; Li, M.; Chan, S.I. Interaction of flavones and their bromoacetyl derivatives with NAD(P)H:quinone acceptor oxidoreductase. Mol. Pharmacol. 1995, 47, 419–424. [Google Scholar] [PubMed]
  54. Medina, J.H.; Viola, H.; Wolfman, C.; Marder, M.; Wasowski, C.; Calvo, D.; Paladini, A.C. Overview—Flavonoids: A new family of benzodiazepine receptor ligands. Neurochem. Res. 1997, 22, 419–425. [Google Scholar] [CrossRef] [PubMed]
  55. Kavvadias, D.; Monschein, V.; Sand, P.; Riederer, P.; Schreier, P. Constituents of sage (Salvia officinalis) with in vitro affinity to human brain benzodiazepine receptor. Planta Med. 2003, 69, 113–117. [Google Scholar] [CrossRef] [PubMed]
  56. Hieble, J.P.; Kolpak, D.C.; McCafferty, G.P.; Ruffolo, R.R., Jr.; Testa, R.; Leonardi, A. Effects of alpha1-adrenoceptor antagonists on agonist and tilt-induced changes in blood pressure: Relationships to uroselectivity. Eur. J. Pharmacol. 1999, 373, 51–62. [Google Scholar] [CrossRef]
  57. Testa, R.; Guarneri, L.; Taddei, C.; Poggesi, E.; Angelico, P.; Sartani, A.; Leonardi, A.; Gofrit, O.N.; Meretyk, S.; Caine, M. Functional antagonistic activity of Rec 15/2739, a novel alpha-1 antagonist selective for the lower urinary tract, on noradrenaline-induced contraction of human prostate and mesenteric artery. J. Pharmacol. Exp. Ther. 1996, 277, 1237–1246. [Google Scholar]
  58. Shin, J.S.; Kim, K.S.; Kim, M.B.; Jeong, J.H.; Kim, B.K. Synthesis and hypoglycemic effect of chrysin derivatives. Bioorg. Med. Chem. Lett. 1999, 9, 869–874. [Google Scholar] [CrossRef]
  59. Hari Babu, T.; Rama Subba Rao, V.; Tiwari, A.K.; Suresh Babu, K.; Srinivas, P.V.; Ali, A.Z.; Madhusudana Rao, J. Synthesis and biological evaluation of novel 8-aminomethylated oroxylin A analogues as alpha-glucosidase inhibitors. Bioorg. Med. Chem. Lett. 2008, 18, 1659–1662. [Google Scholar] [CrossRef]
  60. Kim, J.S.; Kwon, C.S.; Son, K.H. Inhibition of alpha-glucosidase and amylase by luteolin, a flavonoid. Biosci. Biotechnol. Biochem. 2000, 64, 2458–2461. [Google Scholar] [CrossRef]
  61. Kato, A.; Nasu, N.; Takebayashi, K.; Adachi, I.; Minami, Y.; Sanae, F.; Asano, N.; Watson, A.A.; Nash, R.J. Structure-activity relationships of flavonoids as potential inhibitors of glycogen phosphorylase. J. Agric. Food Chem. 2008, 56, 4469–4473. [Google Scholar] [CrossRef]
  62. Okuda, J.; Miwa, I.; Inagaki, K.; Horie, T.; Nakayama, M. Inhibition of aldose reductases from rat and bovine lenses by flavonoids. Biochem. Pharmacol. 1982, 31, 3807–3822. [Google Scholar] [CrossRef]
  63. Porcu, M.; Chiarugi, A. The emerging therapeutic potential of sirtuin-interacting drugs: From cell death to lifespan extension. Trends Pharmacol. Sci. 2005, 26, 94–103. [Google Scholar] [CrossRef] [PubMed]
  64. Marfak, A.; Trouillas, P.; Allais, D.P.; Calliste, C.A.; Cook-Moreau, J.; Duroux, J.L. Reactivity of flavonoids with 1-hydroxyethyl radical: A gamma-radiolysis study. Biochim. Biophys. Acta 2004, 1670, 28–39. [Google Scholar] [CrossRef] [PubMed]
  65. Li, Z.X.; Tian, F.J.; Wu, X.Y.; Zhang, Y.P.; Tian, L.; Shi, S. Anti-arrhythmic action of total flavones of Choerospondias axillaris fructus. Zhongguo Yao Li Xue Bao 1984, 5, 251–254. [Google Scholar] [PubMed]
  66. Li, G.R.; Wang, H.B.; Qin, G.W.; Jin, M.W.; Tang, Q.; Sun, H.Y.; Du, X.L.; Deng, X.L.; Zhang, X.H.; Chen, J.B.; et al. Acacetin, a natural flavone, selectively inhibits human atrial repolarization potassium currents and prevents atrial fibrillation in dogs. Circulation 2008, 117, 2449–2457. [Google Scholar] [CrossRef]
  67. Yang, Y.; Wang, S.; Bao, Y.R.; Li, T.J.; Yang, G.L.; Chang, X.; Meng, X.S. Anti-ulcer effect and potential mechanism of licoflavone by regulating inflammation mediators and amino acid metabolism. J. Ethnopharmacol. 2017, 199, 175–182. [Google Scholar] [CrossRef]
  68. Ninfali, P.; Antonelli, A.; Magnani, M.; Scarpa, E.S. Antiviral Properties of Flavonoids and Delivery Strategies. Nutrients 2020, 12, 2534. [Google Scholar] [CrossRef]
  69. Cushnie, T.P.; Lamb, A.J. Antimicrobial activity of flavonoids. Int. J. Antimicrob. Agents 2005, 26, 343–356. [Google Scholar] [CrossRef]
  70. Aboody, M.S.A.; Mickymaray, S. Anti-Fungal Efficacy and Mechanisms of Flavonoids. Antibiotics 2020, 9, 45. [Google Scholar] [CrossRef]
  71. Borges-Argaez, R.; Balnbury, L.; Flowers, A.; Gimenez-Turba, A.; Ruiz, G.; Waterman, P.G.; Pena-Rodriguez, L.M. Cytotoxic and antiprotozoal activity of flavonoids from Lonchocarpus spp. Phytomedicine 2007, 14, 530–533. [Google Scholar] [CrossRef]
  72. Nunes, A.R.; Vieira, I.G.P.; Queiroz, D.B.; Leal, A.; Maia Morais, S.; Muniz, D.F.; Calixto-Junior, J.T.; Coutinho, H.D.M. Use of Flavonoids and Cinnamates, the Main Photoprotectors with Natural Origin. Adv. Pharmacol. Sci. 2018, 2018, 5341487. [Google Scholar] [CrossRef] [PubMed]
  73. Arct, J.; Pytkowska, K. Flavonoids as components of biologically active cosmeceuticals. Clin. Dermatol. 2008, 26, 347–357. [Google Scholar] [CrossRef] [PubMed]
  74. Ferrell, J.E., Jr.; Chang Sing, P.D.; Loew, G.; King, R.; Mansour, J.M.; Mansour, T.E. Structure/activity studies of flavonoids as inhibitors of cyclic AMP phosphodiesterase and relationship to quantum chemical indices. Mol. Pharmacol. 1979, 16, 556–568. [Google Scholar] [PubMed]
  75. Cao, Z.; Zhang, H.; Cai, X.; Fang, W.; Chai, D.; Wen, Y.; Chen, H.; Chu, F.; Zhang, Y. Luteolin Promotes Cell Apoptosis by Inducing Autophagy in Hepatocellular Carcinoma. Cell. Physiol. Biochem. 2017, 43, 1803–1812. [Google Scholar] [CrossRef]
  76. Lee, D.H.; Kim, C.; Zhang, L.; Lee, Y.J. Role of p53, PUMA, and Bax in wogonin-induced apoptosis in human cancer cells. Biochem. Pharmacol. 2008, 75, 2020–2033. [Google Scholar] [CrossRef]
  77. Cathcart, M.C.; Useckaite, Z.; Drakeford, C.; Semik, V.; Lysaght, J.; Gately, K.; O’Byrne, K.J.; Pidgeon, G.P. Anti-cancer effects of baicalein in non-small cell lung cancer in-vitro and in-vivo. BMC Cancer 2016, 16, 707. [Google Scholar] [CrossRef]
  78. Pan, M.H.; Lai, C.S.; Hsu, P.C.; Wang, Y.J. Acacetin induces apoptosis in human gastric carcinoma cells accompanied by activation of caspase cascades and production of rea.active oxygen species. J. Agric. Food Chem. 2005, 53, 620–630. [Google Scholar] [CrossRef]
  79. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef]
  80. Zhou, Y.; Zheng, J.; Li, Y.; Xu, D.P.; Li, S.; Chen, Y.M.; Li, H.B. Natural Polyphenols for Prevention and Treatment of Cancer. Nutrients 2016, 8, 515. [Google Scholar] [CrossRef]
  81. 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]
  82. Madunic, J.; Madunic, I.V.; Gajski, G.; Popic, J.; Garaj-Vrhovac, V. Apigenin: A dietary flavonoid with diverse anticancer properties. Cancer Lett. 2018, 413, 11–22. [Google Scholar] [CrossRef] [PubMed]
  83. Salmani, J.M.M.; Zhang, X.P.; Jacob, J.A.; Chen, B.A. Apigenin’s anticancer properties and molecular mechanisms of action: Recent advances and future prospectives. Chin. J. Nat. Med. 2017, 15, 321–329. [Google Scholar] [CrossRef]
  84. Kim, H.R.; Park, C.G.; Jung, J.Y. Acacetin (5,7-dihydroxy-4′-methoxyflavone) exhibits in vitro and in vivo anticancer activity through the suppression of NF-kappaB/Akt signaling in prostate cancer cells. Int. J. Mol. Med. 2014, 33, 317–324. [Google Scholar] [CrossRef] [PubMed]
  85. Singh, S.; Gupta, P.; Meena, A.; Luqman, S. Acacetin, a flavone with diverse therapeutic potential in cancer, inflammation, infections and other metabolic disorders. Food Chem. Toxicol. 2020, 145, 111708. [Google Scholar] [CrossRef]
  86. Prasad, N.; Sharma, J.R.; Yadav, U.C.S. Induction of growth cessation by acacetin via beta-catenin pathway and apoptosis by apoptosis inducing factor activation in colorectal carcinoma cells. Mol. Biol. Rep. 2020, 47, 987–1001. [Google Scholar] [CrossRef]
  87. Shim, H.Y.; Park, J.H.; Paik, H.D.; Nah, S.Y.; Kim, D.S.; Han, Y.S. Acacetin-induced apoptosis of human breast cancer MCF-7 cells involves caspase cascade, mitochondria-mediated death signaling and SAPK/JNK1/2-c-Jun activation. Mol. Cells 2007, 24, 95–104. [Google Scholar]
  88. Dou, J.; Wang, Z.; Ma, L.; Peng, B.; Mao, K.; Li, C.; Su, M.; Zhou, C.; Peng, G. Baicalein and baicalin inhibit colon cancer using two distinct fashions of apoptosis and senescence. Oncotarget 2018, 9, 20089–20102. [Google Scholar] [CrossRef]
  89. Liu, H.; Dong, Y.; Gao, Y.; Du, Z.; Wang, Y.; Cheng, P.; Chen, A.; Huang, H. The Fascinating Effects of Baicalein on Cancer: A Review. Int. J. Mol. Sci. 2016, 17, 1681. [Google Scholar] [CrossRef]
  90. Lee, M.N.; Lee, Y.; Wu, D.; Pae, M. Luteolin inhibits NLRP3 inflammasome activation via blocking ASC oligomerization. J. Nutr. Biochem. 2021, 92, 108614. [Google Scholar] [CrossRef]
  91. Imran, M.; Rauf, A.; Abu-Izneid, T.; Nadeem, M.; Shariati, M.A.; Khan, I.A.; Imran, A.; Orhan, I.E.; Rizwan, M.; Atif, M.; et al. Luteolin, a flavonoid, as an anticancer agent: A review. Biomed. Pharmacother. 2019, 112, 108612. [Google Scholar] [CrossRef]
  92. Lin, Y.; Shi, R.; Wang, X.; Shen, H.M. Luteolin, a flav.vonoid with potential for cancer prevention and therapy. Curr. Cancer Drug Targets 2008, 8, 634–646. [Google Scholar] [CrossRef]
  93. Raina, R.; Pramodh, S.; Rais, N.; Haque, S.; Shafarin, J.; Bajbouj, K.; Hamad, M.; Hussain, A. Luteolin inhibits proliferation, triggers apoptosis and modulates Akt/mTOR and MAP kinase pathways in HeLa cells. Oncol. Lett. 2021, 21, 192. [Google Scholar] [CrossRef] [PubMed]
  94. Ko, Y.C.; Choi, H.S.; Liu, R.; Kim, J.H.; Kim, S.L.; Yun, B.S.; Lee, D.S. Inhibitory Effects of Tangeretin, A Citrus Peel-Derived Flavonoid, on Breast Cancer Stem Cell Formation through Suppression of Stat3 Signaling. Molecules 2020, 25, 2599. [Google Scholar] [CrossRef] [PubMed]
  95. Dong, Y.; Cao, A.; Shi, J.; Yin, P.; Wang, L.; Ji, G.; Xie, J.; Wu, D. Tangeretin, a citrus polymethoxyflavonoid, induces apoptosis of human gastric cancer AGS cells through extrinsic and intrinsic signaling pathways. Oncol. Rep. 2014, 31, 1788–1794. [Google Scholar] [CrossRef] [PubMed]
  96. Arafa, E.A.; Shurrab, N.T.; Buabeid, M.A. Therapeutic Implications of a Polymethoxylated Flavone, Tangeretin, in the Management of Cancer via Modulation of Different Molecular Pathways. Adv. Pharmacol. Pharm. Sci. 2021, 2021, 4709818. [Google Scholar] [CrossRef]
  97. Raza, W.; Luqman, S.; Meena, A. Prospects of tangeretin as a modulator of cancer targets/pathways. Pharmacol. Res. 2020, 161, 105202. [Google Scholar] [CrossRef]
  98. Ruibin, J.; Bo, J.; Danying, W.; Chihong, Z.; Jianguo, F.; Linhui, G. Therapy Effects of Wogonin on Ovarian Cancer Cells. Biomed. Res. Int. 2017, 2017, 9381513. [Google Scholar] [CrossRef]
  99. Sharifi-Rad, J.; Herrera-Bravo, J.; Salazar, L.A.; Shaheen, S.; Abdulmajid Ayatollahi, S.; Kobarfard, F.; Imran, M.; Imran, A.; Custodio, L.; Dolores Lopez, M.; et al. The Therapeutic Potential of Wogonin Observed in Preclinical Studies. Evid. Based Complement. Alternat. Med. 2021, 2021, 9935451. [Google Scholar] [CrossRef]
  100. Huynh, D.L.; Sharma, N.; Kumar Singh, A.; Singh Sodhi, S.; Zhang, J.J.; Mongre, R.K.; Ghosh, M.; Kim, N.; Ho Park, Y.; Kee Jeong, D. Anti-tumor activity of wogonin, an extract from Scutellaria baicalensis, through regulating different signaling pathways. Chin. J. Nat. Med. 2017, 15, 15–40. [Google Scholar] [CrossRef]
  101. Tsai, C.F.; Yeh, W.L.; Huang, S.M.; Tan, T.W.; Lu, D.Y. Wogonin induces reactive oxygen species production and cell apoptosis in human glioma cancer cells. Int. J. Mol. Sci. 2012, 13, 9877–9892. [Google Scholar] [CrossRef]
  102. Reed, J.C. Bcl-2 family proteins: Regulators of apoptosis and chemoresistance in hematologic malignancies. Semin. Hematol. 1997, 34 (Suppl. 5), 9–19. [Google Scholar]
  103. Fulda, S.; Debatin, K.M. Extrinsic versus intrinsic apoptosis pathways in anticancer chemotherapy. Oncogene 2006, 25, 4798–4811. [Google Scholar] [CrossRef] [PubMed]
  104. Schneider, P.; Tschopp, J. Apoptosis induced by death receptors. Pharm. Acta Helv. 2000, 74, 281–286. [Google Scholar] [CrossRef]
  105. Annovazzi, L.; Mellai, M.; Caldera, V.; Valente, G.; Tessitore, L.; Schiffer, D. mTOR, S6 and AKT expression in relation to proliferation and apoptosis/autophagy in glioma. Anticancer Res. 2009, 29, 3087–3094. [Google Scholar]
  106. Ersahin, T.; Tuncbag, N.; Cetin-Atalay, R. The PI3K/AKT/mTOR interactive pathway. Mol. Biosyst. 2015, 11, 1946–1954. [Google Scholar] [CrossRef]
  107. Yang, J.; Nie, J.; Ma, X.; Wei, Y.; Peng, Y.; Wei, X. Targeting PI3K in cancer: Mechanisms and advances in clinical trials. Mol. Cancer 2019, 18, 26. [Google Scholar] [CrossRef]
  108. Fresno Vara, J.A.; Casado, E.; de Castro, J.; Cejas, P.; Belda-Iniesta, C.; Gonzalez-Baron, M. PI3K/Akt signalling pathway and cancer. Cancer Treat. Rev. 2004, 30, 193–204. [Google Scholar] [CrossRef]
  109. Testa, J.R.; Bellacosa, A. AKT plays a central role in tumorigenesis. Proc. Natl. Acad. Sci. USA 2001, 98, 10983–10985. [Google Scholar] [CrossRef]
  110. Duan, S.; Huang, W.; Liu, X.; Liu, X.; Chen, N.; Xu, Q.; Hu, Y.; Song, W.; Zhou, J. IMPDH2 promotes colorectal cancer progression through activation of the PI3K/AKT/mTOR and PI3K/AKT/FOXO1 signaling pathways. J. Exp. Clin. Cancer Res. 2018, 37, 304. [Google Scholar] [CrossRef]
  111. Li, H.; Gao, Q.; Guo, L.; Lu, S.H. The PTEN/PI3K/Akt pathway regulates stem-like cells in primary esophageal carcinoma cells. Cancer Biol. Ther. 2011, 11, 950–958. [Google Scholar] [CrossRef]
  112. Li, B.; Cheung, P.Y.; Wang, X.; Tsao, S.W.; Ling, M.T.; Wong, Y.C.; Cheung, A.L. Id-1 activation of PI3K/Akt/NFkappaB signaling pathway and its significance in promoting survival of esophageal cancer cells. Carcinogenesis 2007, 28, 2313–2320. [Google Scholar] [CrossRef] [PubMed]
  113. Seo, B.R.; Min, K.J.; Cho, I.J.; Kim, S.C.; Kwon, T.K. Correction: Curcumin Significantly Enhances Dual PI3K/Akt and mTOR Inhibitor NVP-BEZ235-Induced Apoptosis in Human Renal Carcinoma Caki Cells through Down-Regulation of p53-Dependent Bcl-2 Expression and Inhibition of Mcl-1 Protein Stability. PLoS ONE 2016, 11, e0151886. [Google Scholar]
  114. Pai, S.G.; Carneiro, B.A.; Mota, J.M.; Costa, R.; Leite, C.A.; Barroso-Sousa, R.; Kaplan, J.B.; Chae, Y.K.; Giles, F.J. Wnt/beta-catenin pathway: Modulating anticancer immune response. J. Hematol. Oncol. 2017, 10, 101. [Google Scholar] [CrossRef] [PubMed]
  115. Su, N.; Wang, P.; Li, Y. Role of Wnt/beta-catenin pathway in inducing autophagy and apoptosis in multiple myeloma cells. Oncol. Lett. 2016, 12, 4623–4629. [Google Scholar] [CrossRef]
  116. Valenta, T.; Hausmann, G.; Basler, K. The many faces and functions of beta-catenin. EMBO J. 2012, 31, 2714–2736. [Google Scholar] [CrossRef]
  117. Chen, S.; Guttridge, D.C.; You, Z.; Zhang, Z.; Fribley, A.; Mayo, M.W.; Kitajewski, J.; Wang, C.Y. Wnt-1 signaling inhibits apoptosis by activating beta-catenin/T cell factor-mediated transcription. J. Cell Biol. 2001, 152, 87–96. [Google Scholar] [CrossRef]
  118. de Almeida, G.C.; Oliveira, L.F.S.; Predes, D.; Fokoue, H.H.; Kuster, R.M.; Oliveira, F.L.; Mendes, F.A.; Abreu, J.G. Piperine suppresses the Wnt/beta-catenin pathway and has anti-cancer effects on colorectal cancer cells. Sci. Rep. 2020, 10, 11681. [Google Scholar] [CrossRef]
  119. Trejo-Solis, C.; Escamilla-Ramirez, A.; Jimenez-Farfan, D.; Castillo-Rodriguez, R.A.; Flores-Najera, A.; Cruz-Salgado, A. Crosstalk of the Wnt/beta-Catenin Signaling Pathway in the Induction of Apoptosis on Cancer Cells. Pharmaceuticals 2021, 14, 871. [Google Scholar] [CrossRef]
  120. Bowman, T.; Garcia, R.; Turkson, J.; Jove, R. STATs in oncogenesis. Oncogene 2000, 19, 2474–2488. [Google Scholar] [CrossRef]
  121. Morris, R.; Kershaw, N.J.; Babon, J.J. The molecular details of cytokine signaling via the JAK/STAT pathway. Protein Sci. 2018, 27, 1984–2009. [Google Scholar] [CrossRef]
  122. Juan, W.C.; Ong, S.T. The role of protein phosphorylation in therapy resistance and disease progression in chronic myelogenous leukemia. Prog. Mol. Biol. Transl. Sci. 2012, 106, 107–142. [Google Scholar] [PubMed]
  123. Seif, F.; Khoshmirsafa, M.; Aazami, H.; Mohsenzadegan, M.; Sedighi, G.; Bahar, M. The role of JAK-STAT signaling pathway and its regulators in the fate of T helper cells. Cell Commun. Signal. 2017, 15, 23. [Google Scholar] [CrossRef] [PubMed]
  124. Bose, S.; Banerjee, S.; Mondal, A.; Chakraborty, U.; Pumarol, J.; Croley, C.R.; Bishayee, A. Targeting the JAK/STAT Signaling Pathway Using Phytocompounds for Cancer Prevention and Therapy. Cells 2020, 9, 1451. [Google Scholar] [CrossRef] [PubMed]
  125. Guo, Y.J.; Pan, W.W.; Liu, S.B.; Shen, Z.F.; Xu, Y.; Hu, L.L. ERK/MAPK signalling pathway and tumorigenesis. Exp. Ther. Med. 2020, 19, 1997–2007. [Google Scholar] [CrossRef] [PubMed]
  126. Lee, S.; Rauch, J.; Kolch, W. Targeting MAPK Signaling in Cancer: Mechanisms of Drug Resistance and Sensitivity. Int. J. Mol. Sci. 2020, 21, 1102. [Google Scholar] [CrossRef]
  127. Wada, T.; Penninger, J.M. Mitogen-activated protein kinases in apoptosis regulation. Oncogene 2004, 23, 2838–2849. [Google Scholar] [CrossRef]
  128. Hilger, R.A.; Scheulen, M.E.; Strumberg, D. The Ras-Raf-MEK-ERK pathway in the treatment of cancer. Onkologie 2002, 25, 511–518. [Google Scholar] [CrossRef]
  129. English, J.M.; Cobb, M.H. Pharmacological inhibitors of MAPK pathways. Trends Pharmacol. Sci. 2002, 23, 40–45. [Google Scholar] [CrossRef]
  130. Yue, J.; Lopez, J.M. Understanding MAPK Signaling Pathways in Apoptosis. Int. J. Mol. Sci. 2020, 21, 2346. [Google Scholar] [CrossRef]
  131. Timofeev, O. Editorial: Mutant p53 in Cancer Progression and Personalized Therapeutic Treatments. Front. Oncol. 2021, 11, 740578. [Google Scholar] [CrossRef]
  132. Aubrey, B.J.; Kelly, G.L.; Janic, A.; Herold, M.J.; Strasser, A. How does p53 induce apoptosis and how does this relate to p53-mediated tumour suppression? Cell Death Differ. 2018, 25, 104–113. [Google Scholar] [CrossRef] [PubMed]
  133. Nakano, K.; Vousden, K.H. PUMA, a novel proapop.ptotic gene, is induced by p53. Mol. Cell 2001, 7, 683–694. [Google Scholar] [CrossRef]
  134. Oda, E.; Ohki, R.; Murasawa, H.; Nemoto, J.; Shibue, T.; Yamashita, T.; Tokino, T.; Taniguchi, T.; Tanaka, N. Noxa, a BH3-only member of the Bcl-2 family and candidate mediator of p53-induced apoptosis. Science 2000, 288, 1053–1058. [Google Scholar] [CrossRef] [PubMed]
  135. O’Connor, L.; Strasser, A.; O’Reilly, L.A.; Hausmann, G.; Adams, J.M.; Cory, S.; Huang, D.C. Bim: A novel member of the Bcl-2 family that promotes apoptosis. EMBO J. 1998, 17, 384–395. [Google Scholar] [CrossRef]
  136. 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]
  137. 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.G. Apigenin induces apoptosis via extrinsic pathway, inducing p53 and inhibiting STAT3 and NFkappaB signaling in HER2-overexpressing breast cancer cells. Mol. Cell. Biochem. 2012, 366, 319–334. [Google Scholar] [CrossRef]
  138. 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]
  139. Yang, J.; Pi, C.; Wang, G. Inhibition of PI3K/Akt/mTOR pathway by apigenin induces apoptosis and autophagy in hepatocellular carcinoma cells. Biomed. Pharmacother. 2018, 103, 699–707. [Google Scholar] [CrossRef]
  140. Lee, Y.; Sung, B.; Kang, Y.J.; Kim, D.H.; Jang, J.Y.; Hwang, S.Y.; Kim, M.; Lim, H.S.; Yoon, J.H.; Chung, H.Y.; et al. Apigenin-induced apoptosis is enhanced by inhibition of autophagy formation in HCT116 human colon cancer cells. Int. J. Oncol. 2014, 44, 1599–1606. [Google Scholar] [CrossRef]
  141. Koosha, S.; Alshawsh, M.A.; Looi, C.Y.; Seyedan, A.; Mohamed, Z. An Association Map on the Effect of Flavonoids on the Signaling Pathways in Colorectal Cancer. Int. J. Med. Sci. 2016, 13, 374–385. [Google Scholar] [CrossRef]
  142. Chidambara Murthy, K.N.; Kim, J.; Vikram, A.; Patil, B.S. Differential inhibition of human colon cancer cells by structurally similar flavonoids of citrus. Food Chem. 2012, 132, 27–34. [Google Scholar] [CrossRef]
  143. Maeda, Y.; Takahashi, H.; Nakai, N.; Yanagita, T.; Ando, N.; Okubo, T.; Saito, K.; Shiga, K.; Hirokawa, T.; Hara, M.; et al. Apigenin induces apoptosis by suppressing Bcl-xl and Mcl-1 simultaneously via signal transducer and activator of transcription 3 signaling in colon cancer. Int. J. Oncol. 2018, 52, 1661–1673. [Google Scholar] [CrossRef] [PubMed]
  144. Woo, J.S.; Choo, G.S.; Yoo, E.S.; Kim, S.H.; Lee, J.H.; Han, S.H.; Kim, H.J.; Jung, S.H.; Park, Y.S.; Kim, B.S.; et al. Apigenin induces apoptosis by regulating Akt and MAPK pathways in human melanoma cell A375SM. Mol. Med. Rep. 2020, 22, 4877–4889. [Google Scholar] [CrossRef] [PubMed]
  145. Lu, H.F.; Chie, Y.J.; Yang, M.S.; Lee, C.S.; Fu, J.J.; Yang, J.S.; Tan, T.W.; Wu, S.H.; Ma, Y.S.; Ip, S.W.; et al. Apigenin induces caspase-dependent apoptosis in human lung cancer A549 cells through Bax- and Bcl-2-triggered mitochondrial pathway. Int. J. Oncol. 2010, 36, 1477–1484. [Google Scholar] [PubMed]
  146. 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] [PubMed]
  147. Lee, J.H.; Li, Y.C.; Ip, S.W.; Hsu, S.C.; Chang, N.W.; Tang, N.Y.; Yu, C.S.; Chou, S.T.; Lin, S.S.; Lino, C.C.; et al. The role of Ca2+ in baicalein-induced apoptosis in human breast MDA-MB-231 cancer cells through mitochondria- and caspase-3-dependent pathway. Anticancer Res. 2008, 28, 1701–1711. [Google Scholar]
  148. Yan, W.; Ma, X.; Zhao, X.; Zhang, S. Baicalein induces apoptosis and autophagy of breast cancer cells via inhibiting PI3K/AKT pathway in vivo and vitro. Drug Des. Dev. Ther. 2018, 12, 3961–3972. [Google Scholar] [CrossRef]
  149. Wang, Z.; Jiang, C.; Chen, W.; Zhang, G.; Luo, D.; Cao, Y.; Wu, J.; Ding, Y.; Liu, B. Baicalein induces apoptosis and autophagy via endoplasmic reticulum stress in hepatocellular carcinoma cells. Biomed. Res. Int. 2014, 2014, 732516. [Google Scholar] [CrossRef]
  150. Mu, J.; Liu, T.; Jiang, L.; Wu, X.; Cao, Y.; Li, M.; Dong, Q.; Liu, Y.; Xu, H. The Traditional Chinese Medicine Baicalein Potently Inhibits Gastric Cancer Cells. J. Cancer 2016, 7, 453–461. [Google Scholar] [CrossRef] [Green Version]
  151. Gao, C.; Zhou, Y.; Li, H.; Cong, X.; Jiang, Z.; Wang, X.; Cao, R.; Tian, W. Antitumor effects of baicalin on ovarian cancer cells through induction of cell apoptosis and inhibition of cell migration in vitro. Mol. Med. Rep. 2017, 16, 8729–8734. [Google Scholar] [CrossRef]
  152. 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] [PubMed]
  153. Huang, L.; Jin, K.; Lan, H. Luteolin inhibits cell cycle progression and induces apoptosis of breast cancer cells through downregulation of human telomerase reverse transcriptase. Oncol. Lett. 2019, 17, 3842–3850. [Google Scholar] [CrossRef]
  154. Shi, R.; Huang, Q.; Zhu, X.; Ong, Y.B.; Zhao, B.; Lu, J.; Ong, C.N.; Shen, H.M. Luteolin sensitizes the anticancer effect of cisplatin via c-Jun NH2-terminal kinase-mediated p53 phosphorylation and stabilization. Mol. Cancer Ther. 2007, 6, 1338–1347. [Google Scholar] [CrossRef] [PubMed]
  155. Jiang, Z.Q.; Li, M.H.; Qin, Y.M.; Jiang, H.Y.; Zhang, X.; Wu, M.H. Luteolin Inhibits Tumorigenesis and Induces Apoptosis of Non-Small Cell Lung Cancer Cells via Regulation of MicroRNA-34a-5p. Int. J. Mol. Sci. 2018, 19, 447. [Google Scholar] [CrossRef] [PubMed]
  156. Lu, X.; Li, Y.; Li, X.; Aisa, H.A. Luteolin induces apoptosis in vitro through suppressing the MAPK and PI3K signaling pathways in gastric cancer. Oncol. Lett. 2017, 14, 1993–2000. [Google Scholar] [CrossRef] [PubMed]
  157. Lee, H.J.; Wang, C.J.; Kuo, H.C.; Chou, F.P.; Jean, L.F.; Tseng, T.H. Induction apoptosis of luteolin in human hepatoma HepG2 cells involving mitochondria translocation of Bax/Bak and activation of JNK. Toxicol. Appl. Pharmacol. 2005, 203, 124–131. [Google Scholar] [CrossRef]
  158. Fan, S.; Xu, H.; Liu, H.; Hu, Z.; Xiao, J.; Liu, H. Inhibition of cancer cell growth by Tangeretin flavone in drug-resistant MDA-MB-231 human breast carcinoma cells is facilitated via targeting cell apoptosis, cell cycle phase distribution, cell invasion and activation of numerous Caspases. J. BUON 2019, 24, 1532–1537. [Google Scholar]
  159. Huang, K.F.; Zhang, G.D.; Huang, Y.Q.; Diao, Y. Wogonin induces apoptosis and down-regulates survivin in human breast cancer MCF-7 cells by modulating PI3K-AKT pathway. Int. Immunopharmacol. 2012, 12, 334–341. [Google Scholar] [CrossRef]
  160. Kim, S.J.; Kim, H.J.; Kim, H.R.; Lee, S.H.; Cho, S.D.; Choi, C.S.; Nam, J.S.; Jung, J.Y. Antitumor actions of baicalein and wogonin in HT-29 human colorectal cancer cells. Mol. Med. Rep. 2012, 6, 1443–1449. [Google Scholar] [CrossRef]
  161. He, L.; Lu, N.; Dai, Q.; Zhao, Y.; Zhao, L.; Wang, H.; Li, Z.; You, Q.; Guo, Q. Wogonin induced G1 cell cycle arrest by regulating Wnt/beta-catenin signaling pathway and inactivating CDK8 in human colorectal cancer carcinoma cells. Toxicology 2013, 312, 36–47. [Google Scholar] [CrossRef]
  162. Tan, H.; Li, X.; Yang, W.H.; Kang, Y. A flavone, Wogonin from Scutellaria baicalensis inhibits the proliferation of human colorectal cancer cells by inducing of autophagy, apoptosis and G2/M cell cycle arrest via modulating the PI3K/AKT and STAT3 signalling pathways. J. BUON 2019, 24, 1143–1149. [Google Scholar] [PubMed]
  163. Banik, K.; Khatoon, E.; Harsha, C.; Rana, V.; Parama, D.; Thakur, K.K.; Bishayee, A.; Kunnumakkara, A.B. Wogonin and its analogs for the prevention and treatment of cancer: A systematic review. Phytother. Res. 2022, 36, 1854–1883. [Google Scholar] [CrossRef] [PubMed]
  164. 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]
  165. Shukla, S.; Kanwal, R.; Shankar, E.; Datt, M.; Chance, M.R.; Fu, P.; MacLennan, G.T.; Gupta, S. Apigenin blocks IKKalpha activation and suppresses prostate cancer progression. Oncotarget 2015, 6, 31216–31232. [Google Scholar] [CrossRef]
  166. Amawi, H.; Ashby, C.R., Jr.; Tiwari, A.K. Cancer chemoprevention through dietary flavonoids: What’s limiting? Chin. J. Cancer 2017, 36, 50. [Google Scholar] [CrossRef]
  167. Costea, T.; Hudita, A.; Ciolac, O.A.; Galateanu, B.; Ginghina, O.; Costache, M.; Ganea, C.; Mocanu, M.M. Chemoprevention of Colorectal Cancer by Dietary Compounds. Int. J. Mol. Sci. 2018, 19, 3787. [Google Scholar] [CrossRef]
  168. Yun, S.; Lee, Y.J.; Choi, J.; Kim, N.D.; Han, D.C.; Kwon, B.M. Acacetin Inhibits the Growth of STAT3-Activated DU145 Prostate Cancer Cells by Directly Binding to Signal Transducer and Activator of Transcription 3 (STAT3). Molecules 2021, 26, 6204. [Google Scholar] [CrossRef] [PubMed]
  169. Wang, Y.; Han, E.; Xing, Q.; Yan, J.; Arrington, A.; Wang, C.; Tully, D.; Kowolik, C.M.; Lu, D.M.; Frankel, P.H.; et al. Baicalein upregulates DDIT4 expression which mediates mTOR inhibition and growth inhibition in cancer cells. Cancer Lett. 2015, 358, 170–179. [Google Scholar] [CrossRef]
  170. Peng, Y.; Guo, C.; Yang, Y.; Li, F.; Zhang, Y.; Jiang, B.; Li, Q. Baicalein induces apoptosis of human cervical cancer HeLa cells in vitro. Mol. Med. Rep. 2015, 11, 2129–2134. [Google Scholar] [CrossRef]
  171. Zang, M.D.; Hu, L.; Fan, Z.Y.; Wang, H.X.; Zhu, Z.L.; Cao, S.; Wu, X.Y.; Li, J.F.; Su, L.P.; Li, C.; et al. Luteolin suppresses gastric cancer progression by reversing epithelial-mesenchymal transition via suppression of the Notch signaling pathway. J. Transl. Med. 2017, 15, 52. [Google Scholar] [CrossRef] [Green Version]
  172. Hong, Z.; Cao, X.; Li, N.; Zhang, Y.; Lan, L.; Zhou, Y.; Pan, X.; Shen, L.; Yin, Z.; Luo, L. Luteolin is effective in the non-small cell lung cancer model with L858R/T790M EGF receptor mutation and erlotinib resistance. Br. J. Pharmacol. 2014, 171, 2842–2853. [Google Scholar] [CrossRef]
  173. Wang, Y.; Chen, Y.; Zhang, H.; Chen, J.; Cao, J.; Chen, Q.; Li, X.; Sun, C. Polymethoxyflavones from citrus inhibited gastric cancer cell proliferation through inducing apoptosis by upregulating RARbeta, both in vitro and in vivo. Food Chem. Toxicol. 2020, 146, 111811. [Google Scholar] [CrossRef] [PubMed]
  174. Zhao, Y.; Zhang, L.; Wu, Y.; Dai, Q.; Zhou, Y.; Li, Z.; Yang, L.; Guo, Q.; Lu, N. Selective anti-tumor activity of wogonin targeting the Warburg effect through stablizing p53. Pharmacol. Res. 2018, 135, 49–59. [Google Scholar] [CrossRef]
  175. Batra, P.; Sharma, A.K. Anti-cancer potential of flavonoids: Recent trends and future perspectives. 3 Biotech 2013, 3, 439–459. [Google Scholar] [CrossRef]
  176. Kay, C.D. The future of flavonoid research. Br. J. Nutr. 2010, 104 (Suppl. 3), S91–S95. [Google Scholar] [CrossRef] [PubMed]
  177. Zhai, S.; Senderowicz, A.M.; Sausville, E.A.; Figg, W.D. Flavopiridol, a novel cyclin-dependent kinase inhibitor, in clinical development. Ann. Pharmacother. 2002, 36, 905–911. [Google Scholar] [CrossRef] [PubMed]
  178. Martucciello, S.; Masullo, M.; Cerulli, A.; Piacente, S. Natural Products Targeting ER Stress, and the Functional Link to Mitochondria. Int. J. Mol. Sci. 2020, 21, 1905. [Google Scholar] [CrossRef]
  179. Amodio, G.; Moltedo, O.; Fasano, D.; Zerillo, L.; Oliveti, M.; Di Pietro, P.; Faraonio, R.; Barone, P.; Pellecchia, M.T.; De Rosa, A.; et al. PERK-Mediated Unfolded Protein Response Activation and Oxidative Stress in PARK20 Fibroblasts. Front. Neurosci. 2019, 13, 673. [Google Scholar]
  180. Puthalakath, H.; O’Reilly, L.A.; Gunn, P.; Lee, L.; Kelly, P.N.; Huntington, N.D.; Hughes, P.D.; Michalak, E.M.; McKimm-Breschkin, J.; Motoyama, N.; et al. ER stress triggers apoptosis by activating BH3-only protein Bim. Cell 2007, 129, 1337–1349. [Google Scholar] [CrossRef]
  181. Reimertz, C.; Kogel, D.; Rami, A.; Chittenden, T.; Prehn, J.H. Gene expression during ER stress-induced apoptosis in neurons: Induction of the BH3-only protein Bbc3/PUMA and activation of the mitochondrial apoptosis pathway. J. Cell Biol. 2003, 162, 587–597. [Google Scholar] [CrossRef]
  182. Malhi, H.; Kaufman, R.J. Endoplasmic reticulum stress in liver disease. J. Hepatol. 2011, 54, 795–809. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. The basic structures and subclasses of flavonoids.
Figure 1. The basic structures and subclasses of flavonoids.
Ijms 23 10965 g001
Figure 2. Chemical structures of flavones (A) apigenin, (B) acacetin, (C) baicalein, (D) luteolin, (E) tangeretin, and (F) wogonin.
Figure 2. Chemical structures of flavones (A) apigenin, (B) acacetin, (C) baicalein, (D) luteolin, (E) tangeretin, and (F) wogonin.
Ijms 23 10965 g002
Figure 3. The extrinsic and intrinsic pathways of apoptosis.
Figure 3. The extrinsic and intrinsic pathways of apoptosis.
Ijms 23 10965 g003
Figure 4. Various signaling pathways involved in apoptosis.
Figure 4. Various signaling pathways involved in apoptosis.
Ijms 23 10965 g004
Figure 5. Up- and downregulation of related pathways and proteins in which cancer cells undergo apoptosis by flavones.
Figure 5. Up- and downregulation of related pathways and proteins in which cancer cells undergo apoptosis by flavones.
Ijms 23 10965 g005
Table 1. Biological and pharmacological activities of flavones.
Table 1. Biological and pharmacological activities of flavones.
Biological and Pharmacological Activities of FlavonesReference
  • Protecting cell membrane
[40,41]
  • Antioxidants
[42,43,44]
  • Xanthine oxidase inhibitors
[45]
  • Lipid-lowering agents
[46]
Antiatherogenic agents
-
Nitric oxide synthase inhibitors
-
Cyclo-oxygenase inhibitors
[47,48,49,50]
  • Leukotriene inhibitors
[51]
  • Spasmolytic agents
[52]
  • Prostate hyperplasia therapeutics
[56,57]
  • Phosphodiesterase inhibitors
[74]
  • GABA antagonists
[54,55]
  • NAD(P)H quinone acceptor oxidoreductase inhibitors
[53]
  • Antiarrhythmic agents
[65,66]
  • Anti-ulcer agents
[67]
  • Antivirals
[68]
  • Antibacterials
[69]
  • Antifungals
[70]
Antihyperglycemic
-
α-amylase inhibitors
-
Glycogen phosphorylase inhibitors
-
Aldose reductase inhibitors
Sirtuin activators
[58,59,60,61,62,63]
  • Antiprotozoals
[71]
  • Hepatoprotective agents
[41,64]
  • Photo-protectants
[72]
  • Cosmetic agents
[73]
Anticancer agents
-
Mitosis inhibition
-
Angiogenesis inhibition
-
Protein tyrosine kinase inhibition
-
Ornithine decarboxylase inhibition
-
Aromatase inhibition
Tubulin polymerization inhibitor
[23,41]
Table 2. The six selected flavones with anticancer effects.
Table 2. The six selected flavones with anticancer effects.
FlavoneMolecular FormulaMolecular WeightSourceAnticancer EffectsReference
ApigeninC15H10O5270.24 g/molParsley, grapes, apples, chamomile tea, and red wineBreast, prostate, liver, skin, colorectal, and lung cancers[83]
AcacetinC16H12O5284.26 g/molSafflower, propolis, and Asteraceae plantsBreast, stomach, gastric, and prostate cancers [78,84,85,86,87]
BaicaleinC15H10O5270.24 g/molScutellaria baicalensisBreast, liver, gastric, stomach, and ovarian cancers [88,89]
LuteolinC15H10O6286.24 g/molBroccoli, carrots, celery, cabbage, and parsleyBreast, lung, stomach, liver, and cervical cancers [91,92,93]
TangeretinC20H20O7372.37 g/molCitrus familyBreast, gastric, prostate, and bladder cancers [95,96,97]
WogininC16H12O5284.26 g/molScutellaria baicalensis Georgi (Lamiaceae)Breast, colorectal, lung, and ovarian cancers and glioma[98,100,101]
Table 3. Relationship between flavones and signaling pathways leading to apoptosis in various cancer cells.
Table 3. Relationship between flavones and signaling pathways leading to apoptosis in various cancer cells.
Apigenin
CancerCell LineTreatment
Concentration
Apoptotic Regulation of Cell Signaling PathwayReference
Breast CancerBT-47420, 40, 60, 80, 100 μM
-
Caspase-8 and PARP cleavage
-
STAT3 inhibition
[136]
MCF-710, 20, 40 μM
-
Caspase-8 and PARP cleavage
-
p53-dependent
[137]
Prostate CancerPC-3, DU1455, 10, 20, 40 μM
-
Bcl-2 and Bcl-xL decrease
-
BAX increase
-
Inhibition of XIAP, c-IAP1, c-IAP2, and survivin
[138]
Liver CancerHep G210, 20, 40 μM
-
PI3K/AKT/mTOR inhibition
[139]
Colorectal (Colon) CancerHCT-1166.5, 12.5, 25, 50 μM
-
p53 and p21 increase
[140]
SW48012.5, 25, 50, 100, 200 μM
-
Bcl-2 decrease
-
Caspase-3 and BAX increase
[141,142]
HT29, COLO320, DLD-1, HCT-1165, 15, 50 μM
-
STAT3 phosphorylation inhibition
-
Bcl-xL Mcl-1 decrease
[143]
Human MelanomaA375SM50, 100 μM
-
AKT and MAPK regulation
[144]
Lung CancerA54940, 80, 120, 160 μM
-
Caspase-3, -8, and -9-inducing
-
Caspase-3 and -9 activation
-
Cytochrome c and AIF activation
[145]
NSCLC5, 10, 20, 40, 80, 160 μM
-
TRAIL-inducing (Death Receptors 4 and 5-upregulated)
-
BAD and BAX increase
-
Bcl-2 and Bcl-xL decrease
-
NF-кB and ERK inhibition
[146]
Acacetin
CancerCell LineTreatment
Concentration
Apoptotic Regulation of Cell Signaling PathwayReference
Breast CancerMCF-725, 50, 100, 150, 200 μM
-
Bcl-2 decrease
-
Cytochrome c and AIF activation
-
SAPK/JNK1/2 and c-Jun activation
[87]
Gastric CancerAGS30, 60, 100 μM
-
Caspase activation by ROS generation, mitochondrial-mediated, and Fas activation
[78]
Colorectal (Colon) CancerSW480, HCT-11625 μM
-
β-catenin downregulation
-
BAX increase, Bcl-2 decrease, and AIF-inducing
[86]
Prostate CancerDU14512.5, 25 μM
-
phospho-AKT decrease
-
phospho-GSK-3β decrease
-
p53 increase
-
XIAP and Bcl-2 decrease
-
Activity of phospho-IκB and NF-κB decrease
[84]
Baicalein
CancerCell LineTreatment
Concentration
Apoptotic Regulation of Cell Signaling PathwayReference
Breast CancerMDA-MB-23125, 50, 75, 100 μM
-
Cytochrome c release
-
Caspase-3 activation
-
p53 and ERK/p38MAPK increase
[89,147]
MCF-7, MDA-MB-23110, 20, 40 μM
-
Bcl-2 decrease
-
BAX increase
-
PI3K/AKT inhibition
-
Downregulation of phospho-AKT, phospho-mTOR, NF-кB, and phospho-IкB
[147,148]
Liver CancerHCC25, 50, 100, 200 μM
-
BAX increase
-
Bcl-2 decrease
-
Cleaved caspase-3 and -9 and PARP-inducing
-
JNK activation
[149]
Gastric CancerSGC-790115, 30, 60 μM
-
Mitochondrial-mediated
-
Bcl-2 decrease
-
BAX increase
[150]
Colorectal (Colon) CancerHCT-116, SW48010, 20, 50 μM
-
MAPK/ERK-mediated
[88]
Ovarian CancerA278020, 40, 80, 160 μM
-
Bcl-2 decrease
-
Caspase-3 and -9 activation
[151]
Luteolin
CancerCell LineTreatment
Concentration
Apoptotic Regulation of Cell Signaling PathwayReference
Breast CancerHs578T, MCF-7, MDA-MB-231 12.5, 25, 50, 100 μM
-
PI3K/AKT inhibition
-
FOXO3a activation
[152]
MDA-MB-231 10, 30 μM
-
Downregulation of hTERT expression by NF-κB inhibitor α and c-Myc inhibition
[153]
Lung CancerSCLC20 or 40 μM
-
Increased anticancer effect of cisplatin through jnk activation
[154]
NSCLC10, 20, 40 μM
-
Regulation of microRNA-34a-5p
[155]
Gastric CancerBGC-82320, 40, 60 μM
-
Caspase-3 and -9 increase
-
Cytochrome c increase
-
BAX increase
-
Bcl-2 decrease
-
Inhibition of MAPK and PI3K
[156]
Liver CancerSMMC-772125, 50, 100 μM
-
Caspase-8 increase
-
Bcl-2 decrease
[75]
Hep-G240, 80 μM
-
BAX/BAK mitochondrial translocation
-
JNK activation
[157]
Cervical CancerHeLa5, 10, 20 μM
-
BAX, BAD BID, AFAF1, TRADD, FAS, and FADD increase
-
Caspase-3 and -9 increase
-
Bcl-2 and Mcl-1 decrease
-
Inhibition of AKT and MAPK
[93]
Tangeretin
CancerCell LineTreatment
Concentration
Apoptotic Regulation of cell signaling pathwayReference
Breast CancerMDA-MB-2314.5, 9, 18 μM
-
BAX increase
-
Caspase-3 and-8 increase
-
Bcl-2 decrease
[158]
Gastric CancerAGS10, 30, 60 μM
-
Caspase-3, -8 and -9 increase
-
BAX, tBID, p53 increase
-
Fas/Fas L inducing
[95]
Prostate CancerPC-3, LNCaP25, 50, 75, 100 μM
-
Cleaved caspase-3 and -9 increase
-
EMT inhibition
[97]
PC-325, 50, 100 μM
-
BAX increase
-
Bcl family decrease
DU14525, 50, 100 μM
-
Regulation of the androgen receptor (AR)-PI3K/AKT/mTOR-Notch
Bladder CancerBFTC-90520, 40, 60 μM
-
Cytochrome c and AIF release
-
Cleaved caspase-3, -9, pro-caspase-3 and -9 regulation
[97]
Wogonin
CancerCell LineTreatment
Concentration
Apoptotic Regulation of Cell Signaling PathwayReference
Breast CancerMCF-730, 60, 90 μM
-
Caspase-3, -8, and -9 increase
-
BAX and p53 increase
-
Bcl-2 and survivin decrease
-
Inhibition of PI3K/AKT/survivin
-
ERK activation
[159]
Colorectal (Colon) CancerHT-2925, 50, 100 μM
-
Bcl-2 decrease
-
BAX increase
-
PI3K/AKT-mediated
[160]
HCT11610, 20, 40 μM
-
Inhibition of the β-catenin-dependent Wnt
[161]
SW4804, 8, 16 μM
-
BAX increase
-
Caspase-3, -8, and -9 increase
-
Inhibition of PI3K/AKT/STAT-3
[162]
Lung CancerA5495, 10, 20 μM or 25, 30, 50 μM
-
Caspase-3, -8, and -9 and PARP increase
-
XIAP, c-FLIP, cIAP-1, -2, AKT, and STAT-3 decrease
[163]
Ovarian CancerA278010, 20, 30 μM
-
BAX and p53 increase
-
AKT and Bcl-2 decrease
[98]
GlioblastomaU251, U874, 8, 16, 24 μM
-
Caspase-3, -8, and -9 and PARP increase
-
Generation of ROS
-
Activation of ER stress
[101]
Table 4. Anticancer effects on flavones and in vivo mouse models.
Table 4. Anticancer effects on flavones and in vivo mouse models.
FlavonesType of CancerMouse Model and DosagesIn Vivo FunctionReference
ApigeninColorectal cancerAOM-injected SD rat Suppressed colorectal cancer by reducing
ACF and increasing apoptosis
[141]
Prostate cancerTRAMP mice (20 and 50 μg/mouse/day, gavage)Inhibited the formation of prostate carcinoma by regulating the PI3K/AKT/FOXO pathway[164,165]
Athymic nude mouse
(20 and 50 μg/mouse/day, oral)
Tumor volume was reduced; XIAP, survivin, and HDAC1 were downregulated; and BAX was increased[138]
AcacetinProstate cancerBALB/C nude mouse
(50 mg/kg, intraperitoneal injection 5 days per week for 30 days)
Inhibited tumor growth through STAT3 regulation[168]
BaicaleinLung cancerBALB/C nude mouse
(1 and 3 mg/kg, intratumoral injection twice weekly)
Inhibition of tumor growth, genes that induced apoptosis, ITGB3 and TNFRSF25 upregulated[77]
Breast cancerSCID-Beige mice (20 mg/kg, intraperitoneal injection for 5 days)mTOR inhibition prevented tumor growth[169]
LuteolinGastric cancerBALB/C nude mouse (10 mg/kg, intraperitoneal injection 6 times)Inhibition of Notch1 and β-catenin[171]
Lung cancerBALB/C nude mouse (10 and 30 mg/kg, intraperitoneal injection daily for 15 days)Tumor suppression and tumor weight were reduced by inhibition of PI3K/AKT/mTOR[172]
TangeretinBreast cancerNude mouse (2.5 mg/kg, intraperitoneal injection once a week 4 times)Inhibition of tumor growth[94]
Gastric cancerBALB/C nude mouse (5, 25, 125 mg kg−1·bw−1·day−1)Inhibition of tumor growth[173]
WogoninOvarian cancerAthymic BALB/C nude mouse (20, 40, 80 mg/kg, intraperitoneal injection every 3 days)Tumor volume and weight were reduced[174]
Colorectal cancerAthymic BALB/C nude mouse (20, 40, 80 mg/kg, intraperitoneal injection every 3 days)Tumor volume and weight were reduced[174]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jeong, S.H.; Kim, H.H.; Ha, S.E.; Park, M.Y.; Bhosale, P.B.; Abusaliya, A.; Park, K.I.; Heo, J.D.; Kim, H.W.; Kim, G.S. Flavones: Six Selected Flavones and Their Related Signaling Pathways That Induce Apoptosis in Cancer. Int. J. Mol. Sci. 2022, 23, 10965. https://doi.org/10.3390/ijms231810965

AMA Style

Jeong SH, Kim HH, Ha SE, Park MY, Bhosale PB, Abusaliya A, Park KI, Heo JD, Kim HW, Kim GS. Flavones: Six Selected Flavones and Their Related Signaling Pathways That Induce Apoptosis in Cancer. International Journal of Molecular Sciences. 2022; 23(18):10965. https://doi.org/10.3390/ijms231810965

Chicago/Turabian Style

Jeong, Se Hyo, Hun Hwan Kim, Sang Eun Ha, Min Young Park, Pritam Bhagwan Bhosale, Abuyaseer Abusaliya, Kwang Il Park, Jeong Doo Heo, Hyun Wook Kim, and Gon Sup Kim. 2022. "Flavones: Six Selected Flavones and Their Related Signaling Pathways That Induce Apoptosis in Cancer" International Journal of Molecular Sciences 23, no. 18: 10965. https://doi.org/10.3390/ijms231810965

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